1655
30
The Ocean
Coordinating Lead Authors:
Ove Hoegh-Guldberg (Australia), Rongshuo Cai (China)
Lead Authors:
Elvira S. Poloczanska (Australia), Peter G. Brewer (USA), Svein Sundby (Norway), Karim Hilmi
(Morocco), Victoria J. Fabry (USA), Sukgeun Jung (Republic of Korea)
Contributing Authors:
William Skirving (USA), Dáithí Stone (Canada/South Africa/USA), Michael T. Burrows (UK),
Johann Bell (New Caledonia), Long Cao (China), Simon Donner (Canada), C. Mark Eakin (USA),
Arne Eide (Norway), Benjamin Halpern (USA), Charles R. McClain (USA), Mary I. O’Connor
(Canada), Camille Parmesan (USA), R. Ian Perry (Canada), Anthony J. Richardson (Australia),
Christopher J. Brown (Australia), David Schoeman (Australia), Sergio Signorini (USA),
William Sydeman (USA), Rui Zhang (China), Ruben van Hooidonk (USA), Stewart M. McKinnell
(PICES/Canada)
Review Editors:
Carol Turley (UK), Ly Omar (Senegal)
Volunteer Chapter Scientists:
Jo Davy (New Zealand), Sarah Lee (USA)
This chapter should be cited as:
Hoegh-Guldberg
, O., R. Cai, E.S. Poloczanska, P.G. Brewer, S. Sundby, K. Hilmi, V.J. Fabry, and S. Jung, 2014: The
Ocean. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution
of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change
[Barros, V.R., C.B. Field, D.J. Dokken, M.D. Mastrandrea, K.J. Mach, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada,
R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)].
Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1655-1731.
30
1656
Executive Summary.......................................................................................................................................................... 1658
30.1. Introduction .......................................................................................................................................................... 1662
30.1.1. Major Sub-regions within the Ocean .............................................................................................................................................. 1662
30.1.2. Detection and Attribution of Climate Change and Ocean Acidification in Ocean Sub-regions ........................................................ 1662
30.2. Major Conclusions from Previous Assessments .................................................................................................... 1662
30.3. Regional Changes and Projections of Future Ocean Conditions .......................................................................... 1664
30.3.1. Physical Changes ............................................................................................................................................................................ 1664
30.3.1.1. Heat Content and Temperature ....................................................................................................................................... 1664
30.3.1.2. Sea Level ......................................................................................................................................................................... 1668
30.3.1.3. Ocean Circulation, Surface Wind, and Waves ................................................................................................................... 1671
30.3.1.4. Solar Insolation and Clouds ............................................................................................................................................. 1671
30.3.1.5. Storm Systems ................................................................................................................................................................. 1671
30.3.1.6. Thermal Stratification ...................................................................................................................................................... 1672
30.3.2. Chemical Changes .......................................................................................................................................................................... 1673
30.3.2.1. Surface Salinity ................................................................................................................................................................ 1673
30.3.2.2. Ocean Acidification ......................................................................................................................................................... 1673
30.3.2.3. Oxygen Concentration ..................................................................................................................................................... 1675
30.4. Global Patterns in the Response of Marine Organisms to Climate Change and Ocean Acidification .................. 1677
30.5. Regional Impacts, Risks, and Vulnerabilities: Present and Future ........................................................................ 1677
30.5.1. High-Latitude Spring Bloom Systems .............................................................................................................................................. 1677
30.5.1.1. Observed Changes and Potential Impacts ....................................................................................................................... 1678
30.5.1.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1681
30.5.2. Equatorial Upwelling Systems ........................................................................................................................................................ 1681
30.5.2.1. Observed Changes and Potential Impacts ....................................................................................................................... 1682
30.5.2.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1682
30.5.3. Semi-Enclosed Seas ........................................................................................................................................................................ 1683
30.5.3.1. Observed Changes and Potential Impacts ....................................................................................................................... 1683
30.5.3.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1685
30.5.4. Coastal Boundary Systems .............................................................................................................................................................. 1686
30.5.4.1. Observed Changes and Potential Impacts ....................................................................................................................... 1686
30.5.4.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1688
30.5.5. Eastern Boundary Upwelling Ecosystems ....................................................................................................................................... 1690
30.5.5.1. Observed Changes and Potential Impacts ....................................................................................................................... 1690
30.5.5.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1693
Table of Contents
1657
The Ocean Chapter 30
30
30.5.6. Subtropical Gyres ............................................................................................................................................................................ 1693
30.5.6.1. Observed Changes and Potential Impacts ....................................................................................................................... 1693
30.5.6.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1696
30.5.7. Deep Sea (>1000 m) ....................................................................................................................................................................... 1697
30.5.7.1. Observed Changes and Potential Impacts ....................................................................................................................... 1697
30.5.7.2. Key Risks and Vulnerabilities ........................................................................................................................................... 1698
30.5.8. Detection and Attribution of Climate Change Impacts with Confidence Levels .............................................................................. 1698
30.6. Sectoral Impacts, Adaptation, and Mitigation Responses .................................................................................... 1698
30.6.1. Natural Ecosystem Services ............................................................................................................................................................ 1699
30.6.2. Economic Sectors ............................................................................................................................................................................ 1701
30.6.2.1. Fisheries and Aquaculture ............................................................................................................................................... 1701
30.6.2.2. Tourism ............................................................................................................................................................................ 1704
30.6.2.3. Shipping .......................................................................................................................................................................... 1705
30.6.2.4. Offshore Energy and Mineral Resource Extraction and Supply ........................................................................................ 1705
30.6.3. Human Health ................................................................................................................................................................................ 1705
30.6.4. Ocean-Based Mitigation ................................................................................................................................................................. 1705
30.6.4.1. Deep Sea Carbon Sequestration ...................................................................................................................................... 1705
30.6.4.2. Offshore Renewable Energy ............................................................................................................................................ 1706
30.6.5. Maritime Security and Related Operations ..................................................................................................................................... 1706
30.7. Synthesis and Conclusions .................................................................................................................................... 1706
30.7.1. Key Vulnerabilities and Risks .......................................................................................................................................................... 1708
30.7.2. Global Frameworks for Decision Making ........................................................................................................................................ 1711
30.7.3. Emerging Issues, Data Gaps, and Research Needs .......................................................................................................................... 1713
30.7.3.1. Changing Variability and Marine Impacts ........................................................................................................................ 1713
30.7.3.2. Surface Wind, Storms, and Upwelling .............................................................................................................................. 1713
30.7.3.3. Declining Oxygen Concentrations ................................................................................................................................... 1714
30.7.3.4. Ocean Acidification ......................................................................................................................................................... 1714
30.7.3.5. Net Primary Productivity .................................................................................................................................................. 1714
30.7.3.6. Movement of Marine Organisms and Ecosystems ........................................................................................................... 1714
30.7.3.7. Understanding Cumulative and Synergistic Impacts ........................................................................................................ 1714
30.7.3.8. Reorganization of Ecosystems and Food Webs ................................................................................................................ 1714
30.7.3.9. Socio-ecological Resilience .............................................................................................................................................. 1715
References ....................................................................................................................................................................... 1715
Frequently Asked Questions
30.1: Can we reverse the impacts of climate change on the Ocean? ....................................................................................................... 1675
30.2: Does slower warming in the Ocean mean less impact on plants and animals? .............................................................................. 1677
30.3: How will marine primary productivity change with ocean warming and acidification? .................................................................. 1682
30.4: Will climate change increase the number of “dead zones” in the Ocean? ..................................................................................... 1693
30.5: How can we use non-climate factors to manage climate change impacts on the Ocean? .............................................................. 1710
1658
Chapter 30 The Ocean
30
Executive Summary
The Ocean plays a central role in Earth’s climate and has absorbed 93% of the extra energy from the enhanced greenhouse effect
and approximately 30% of anthropogenic carbon dioxide (CO
2
) from the atmosphere. Regional responses are addressed here by
dividing the Ocean into seven sub-regions: High-Latitude Spring Bloom Systems (HLSBS), Eastern Boundary Upwelling Ecosystems (EBUE),
Coastal Boundary Systems (CBS), Equatorial Upwelling Systems (EUS), Subtropical Gyres (STG), Semi-Enclosed Seas (SES), and the Deep Sea
(DS; >1000 m). An eighth region, Polar Seas, is dealt with by Chapter 28. {Figure 30-1; WGI AR5 6.3.1; WGI AR5 Boxes 3.1, 3.8}
Global average sea surface temperatures have increased since both the beginning of the 20th century and the 1950s (certain).
The average sea surface temperature (SST) of the Indian, Atlantic, and Pacific Oceans has increased by 0.65°C, 0.41°C, and 0.31°C,
respectively, over the period 1950–2009 (very likely, p-value ≤ 0.05).
Changes in the surface temperatures of the ocean basins are
consistent with temperature trends simulated by ocean-atmosphere models with anthropogenic greenhouse gas (GHG) forcing over the past
century (high confidence). Sub-regions within the Ocean also show robust evidence of change, with the influence of long-term patterns of variability
(e.g., Pacific Decadal Oscillation (PDO); Atlantic Multi-decadal Oscillation (AMO)) contributing to variability at regional scales, and making
changes due to climate change harder to distinguish and attribute. {30.3.1; Figure 30-2e-g; Table 30-1; WGI AR5 2.4.2-3, 3.2, 10.4.1, 14}
Uptake of CO
2
has decreased ocean pH (approximately 0.1 unit over 100 years), fundamentally changing ocean carbonate chemistry
in all ocean sub-regions, particularly at high latitudes (high confidence). The current rate of ocean acidification is unprecedented within
the last 65 Ma (high confidence), if not the last 300 Ma (medium confidence). Warming temperatures, and declining pH and carbonate ion
concentrations, represent risks to the productivity of fisheries and aquaculture, and the security of regional livelihoods given the direct and
indirect effects of these variables on physiological processes (e.g., skeleton formation, gas exchange, reproduction, growth, and neural function)
and ecosystem processes (e.g., primary productivity, reef building and erosion) (high confidence). {6.1.2, 6.2-3, 30.3.2, 30.6; WGI AR5 3.8.2;
WGI AR5 Boxes 3.2, 5.3.1}
Regional changes observed in winds, surface salinity, stratification, ocean currents, nutrient availability, and oxygen depth profile
in many regions may be a result of anthropogenic GHG emissions (low to medium confidence). Marine organisms and ecosystems are
likely to change in response to these regional changes, although evidence is limited and responses uncertain. {6.2-3, 30.3, 30.5; WGI AR5 2.7,
3.3-8, 10.4.2, 10.4.4}
Most, if not all, of the Ocean will continue to warm and acidify, although the rates will vary regionally (high confidence). Differences
between Representative Concentration Pathways (RCPs) are very likely to be minimal until 2040 (high confidence). Projected temperatures of
the surface layers of the Ocean, however, diverge as the 21st century unfolds and will be 1°C to 3°C higher by 2100 under RCP8.5 than RCP2.6
across most ocean sub-regions. The projected changes in ocean temperature pose serious risks and vulnerabilities to ocean ecosystems and
dependent human communities (robust evidence, high agreement; high confidence). {6.5, 30.3.1-2, 30.7.1; Figure 30-2e-g; Table 30-3; WGI AR5
11.3.3, 12.4.7; WGI AR5 Box 1.1}
Rapid changes in physical and chemical conditions within ocean sub-regions have already affected the distribution and abundance
of marine organisms and ecosystems.
Responses of species and ecosystems to climate change have been observed from every ocean sub-
region (high confidence). Marine organisms are moving to higher latitudes, consistent with warming trends (high confidence), with fish and
zooplankton migrating at the fastest rates, particularly in HLSBS regions. Changes to sea temperature have also altered the phenology, or timing
of key life-history events such as plankton blooms, and migratory patterns and spawning in fish and invertebrates, over recent decades
(medium confidence). There is medium to high agreement that these changes pose significant uncertainties and risks to fisheries, aquaculture,
and other coastal activities. Ocean acidification maybe driving similar changes (low confidence), although there is limited evidence and low
agreement at present. The associated risks will intensify as ocean warming and acidification continue. {6.3-4, 30.4-5; Table 30-3; Box CC-MB}
Regional risks and vulnerabilities to ocean warming and acidification can be compounded by non-climate related stressors such
as pollution, nutrient runoff from land, and over-exploitation of marine resources, as well as natural climate variability (high
confidence).
These influences confound the detection and attribution of the impacts of climate change and ocean acidification on ecosystems
1659
The Ocean Chapter 30
30
yet may also represent opportunities for reducing risks through management strategies aimed at reducing their influence, especially in CBS,
SES, and HLSBS. {5.3.4, 18.3.3-4, 30.1.2, 30.5-6}
Recent changes to wind and ocean mixing within the highly productive HLSBS, EBUE, and EUS are likely to influence energy
transfer to higher trophic levels and microbial processes. There is, however, limited evidence and low agreement on the direction and
magnitude of these changes and their relationship to ocean warming and acidification (low confidence). In cases where Net Primary Productivity
(NPP) increases or is not consumed (e.g., Benguela EBUE, low confidence), the increased transfer of organic carbon to deep regions can stimulate
microbial respiration and reduce O
2
levels (medium confidence). Oxygen concentrations are also declining in the tropical Pacific, Atlantic, and
Indian Oceans (particularly EUS) due to reduced O
2
solubility at higher temperatures, and changes in ocean ventilation and circulation. {6.3.3,
30.3, 30.5.1-2, 30.5.5; Box CC-PP; WGI AR5 3.8.3}
Global warming will result in more frequent extreme events and greater associated risks to ocean ecosystems (high confidence).
In some cases (e.g., mass coral bleaching and mortality), projected increases will eliminate ecosystems, and increase risks and vulnerabilities to
coastal livelihoods and food security (e.g., CBS in Southeast Asia; SES, CBS, and STG in the Indo-Pacific) (medium to high confidence). Reducing
stressors not related to climate change represents an opportunity to strengthen the ecological resilience within these regions, which may help
them survive some projected changes in ocean temperature and chemistry. {5.4, 30.5.3-4, 30.5.6, 30.6.1; Figure 30-4; Box CC-CR; IPCC, 2012}
The highly productive HLSBS in the Northeastern Atlantic has changed in response to warming (medium evidence, high agreement),
with a range of consequences for fisheries. These ecosystems are responding to recent warming, with the greatest changes being observed
since the late 1970s in the phenology, distribution, and abundance of plankton assemblages, and the reorganization of fish assemblages (high
confidence). There is medium confidence that these changes will have both positive and negative implications depending on the particular
HLSBS fishery and the time frame. {6.4.1.1, 6.5.3, 30.5.1, 30.6.2.1; Boxes CC-MB, 6-1}
EUS, which support highly productive fisheries off equatorial Africa and South America, have warmed over the past 60 years
(Pacific EUS: 0.43°C, Atlantic EUS: 0.54°C; very likely, p-value ≤ 0.05). Although warming is consistent with changes in upwelling intensity,
there is low confidence in our understanding of how EUS will change, especially in how El Niño-Southern Oscillation (ENSO) and other patterns
of variability will interact in a warmer world. The risk, however, of changes to upwelling increases with average global temperature, posing
significant uncertainties for dependent ecosystems, communities, and fisheries. {30.5.2; WGI AR5 14.4}
The surface waters of the SES show significant warming from 1982 and most CBS show significant warming since 1950. Warming
of the Mediterranean has led to the recent spread of tropical species invading from the Atlantic and Indian Oceans. Projected warming increases
the risk of greater thermal stratification in some regions, which can lead to reduced O
2
ventilation and the formation of additional hypoxic
zones, especially in the Baltic and Black Seas (medium confidence). In some CBS, such as the East China Sea and Gulf of Mexico, these changes
are further influenced by the contribution of nutrients from coastal pollution contributing to the expansion of hypoxic (low O
2
) zones. These
changes are likely to influence regional ecosystems as well as dependent industries such as fisheries and tourism, although there is low confidence
in the understanding of potential changes and impacts. {5.3.4.3, 30.5.3-4; Table 30-1}
Coral reefs within CBS, SES, and STG are rapidly declining as a result of local stressors (i.e., coastal pollution, overexploitation)
and climate change (high confidence).
Elevated sea temperatures drive impacts such as mass coral bleaching and mortality (very high
confidence), with an analysis of the Coupled Model Intercomparison Project Phase 5 (CMIP5) ensemble projecting the loss of coral reefs from
most sites globally by 2050 under mid to high rates of ocean warming (very likely). {29.3.1.2, 30.5.3-4, 30.5.6; Figure 30-10; Box CC-CR}
The productive EBUE and EUS involve upwelling waters that are naturally high in CO
2
concentrations and low in pH, and hence
are potentially vulnerable to ocean warming and acidification (medium confidence).
There is limited evidence and low agreement as
to how upwelling systems are likely to change (low confidence). Declining O
2
and shoaling of the aragonite saturation horizon through ocean
acidification increase the risk of upwelling water being low in pH and O
2
, with impacts on coastal ecosystems and fisheries, as has been seen
already (e.g., California Current EBUE). These risks and uncertainties are likely to involve significant challenges for fisheries and associated
1660
Chapter 30 The Ocean
30
livelihoods along the west coasts of South America, Africa, and North America (low to medium confidence). {22.3.2.3, 30.3.2.2, 30.5.2, 30.5.5;
Boxes CC-UP, CC-PP}
Chlorophyll concentrations measured by satellites have decreased in the STG of the North Pacific, Indian, and North Atlantic
Oceans by 9%, 12%, and 11%, respectively, over and above the inherent seasonal and interannual variability from 1998 to 2010
(high confidence; p-value ≤ 0.05).
Significant warming over this period has resulted in increased water column stratification, reduced mixed
layer depth, and possibly decreases in nutrient availability and ecosystem productivity (limited evidence, medium agreement). The short time
frame of these studies against well-established patterns of long-term variability leads to the conclusion that these changes are about as likely
as not due to climate change. {6.3.4, 30.5.6; Table 30-1; Box CC-PP; WGI AR5 3.8.4}
The world’s most abundant yet difficult to access habitat, the DS, is changing (limited evidence, medium agreement), with
warming between 700 and 2000 m from 1957 to 2010 likely to involve a significant anthropogenic signal (medium confidence).
Decreased primary productivity of surface waters (e.g., STG) is likely to reduce the availability of organic carbon to DS ecosystems. Understanding
of the risks of climate change and ocean acidification to the DS is important given the size of the DS region but is limited (low confidence).
{30.5.7; Figure 30-2; WGI AR5 3.2.4; WGI AR5 Figures 3.2, 3.9}
Changes to surface wind and waves, sea level, and storm intensity will increase the vulnerability of ocean-based industries such
as shipping, energy, and mineral extraction (medium confidence).
Risks to equipment and people may be reduced through the design
and use of ocean-based infrastructure, together with the evolution of policy (medium agreement). Risks and uncertainties will increase with
further climate change. New opportunities as well as risks for shipping, energy, and mineral extraction, and international issues over access
and vulnerability, may accompany warming waters, particularly at high latitudes. {10.2.2, 10.4.4, 28.2.6, 28.3.4, 30.3.1, 30.6.2; IPCC, 2012}
Changes to ocean temperature, chemistry, and other factors are generating new challenges for fisheries, as well as benefits (high
agreement). Climate change is a risk to the sustainability of capture fisheries and aquaculture development, adding to the threats of over-
fishing and other non-climate stressors. In EUS and STG, shifts in the distribution and abundance of large pelagic fish stocks will have the
potential to create “winners” and “losers” among island nations and economies. There has been a boost in fish stocks of high-latitude fisheries
in the HLSBS of the North Pacific and North Atlantic, partly as a result of 30 years of increase in temperature. This is very likely to continue,
although some fish stocks will eventually decline. A number of practical adaptation options and supporting international policies can minimize
the risks and maximize the opportunities. {7.4.2, 7.5.1.1.2, 29.4, 30.6-7}
Adaptation strategies for ocean regions beyond coastal waters are generally poorly developed but will benefit from international
legislation and expert networks, as well as marine spatial planning (high agreement).
Fisheries and aquaculture industries with high
technology and/or large investments, as well as marine shipping and oil and gas industries, have high capacities for adaptation due to greater
development of environmental monitoring, modeling, and resource assessments. For smaller scale fisheries and developing nations, building
social resilience, alternative livelihoods, and occupational flexibility represent important strategies for reducing the vulnerability of ocean-
dependent human communities. Building strategies that include climate forecasting and early-warning systems can reduce impacts of warming
and ocean acidification in the short term. Overall, there is a strong need to develop ecosystem-based monitoring and adaptation strategies to
mitigate rapidly growing risks and uncertainties to the coastal and oceanic industries, communities, and nations (high agreement). {7.5.1.1,
30.6}
Significant opportunity exists within the Ocean and its sub-regions for reducing the CO
2
flux to the atmosphere (limited evidence,
medium agreement).
Ecosystems such as mangroves, seagrass, and salt marsh offer important carbon storage and sequestration opportunities
(e.g., Blue Carbon; limited evidence, medium agreement). Blue Carbon strategies can also be justified in terms of the ecosystem services provided
by coastal vegetated habitats such as protection against coastal erosion and storm damage, and maintenance of habitats for fisheries species.
Sequestration of anthropogenic CO
2
into deep ocean areas still faces considerable hurdles with respect to the expense, legality, and vulnerability
of storage sites and infrastructure. There are also significant opportunities with the Ocean for the development of offshore renewable energy
such as wind and tidal power. {5.5.7, 30.6.1, 30.6.4}
1661
The Ocean Chapter 30
30
International frameworks for collaboration and decision making are critically important for coordinating policy that will enable
mitigation and adaptation by the Ocean sectors to global climate change (e.g., United Nations Convention on the Law of the Sea
(UNCLOS)).
These international frameworks offer an opportunity to solve problems collectively, including improving fisheries management
across national borders (e.g., reducing illegal, unreported, and unregulated (IUU) fishing), responding to extreme events, and strengthening
international food security. Given the importance of the Ocean to all countries, there is a need for the international community to progress
rapidly to a “whole of ocean” strategy for responding to the risks and challenges posed by anthropogenic ocean warming and acidification.
{30.7.2}
1662
Chapter 30 The Ocean
30
30.1. Introduction
The Ocean exerts a profound influence as part of the Earth, interacting
with its atmosphere, cryosphere, land, and biosphere to produce planetary
c
onditions. It also directly influences human welfare through the provision
and transport of food and resources, as well as by providing cultural
and economic benefits. The Ocean also contributes to human welfare
indirectly through the regulation of atmospheric gas content and the
distribution of heat and water across the planet. This chapter examines
the extent to which regional changes to the Ocean can be accurately
detected and attributed to anthropogenic climate change and ocean
acidification, building on the conclusions of Chapter 6, which focuses
on the marine physiological and ecological responses to climate change
and ocean acidification. Detailed assessment of the role of recent physical
and chemical changes within the Ocean to anthropogenic climate change
is provided in WGI AR5 (particularly Chapters 2, 3, 13, and 14). In this
chapter, impacts, risks, and vulnerabilities associated with climate change
and ocean acidification are assessed for seven ocean sub-regions, and
the expected consequences and adaptation options for key ocean-based
sectors are discussed. Polar oceans (defined by the presence of sea ice
in the north and by the Polar Front in the south) are considered in
Chapter 28.
Given that climate change affects coastal and low-lying sub-regions of
multiple nations, detailed discussion of potential risks and consequences
for these regions occurs in the relevant chapters of this report (e.g.,
Chapters 5 and 29, as well as other regional sections).
30.1.1. Major Sub-regions within the Ocean
The Ocean represents a vast region that stretches from the high tide
mark to the deepest oceanic trench (11,030 m) and occupies 71% of
the Earth’s surface. The total volume of the Ocean is approximately
1.3 billion km
3
, with approximately 72% of this volume being below
1000 m (Deep Sea (DS); Section 30.5.7). There are considerable challenges
in assessing the regional impacts of climate change on the Ocean.
Devising an appropriate structure to explore the influence of climate
change across the entire Ocean region and the broad diversity of life
forms and habitats is challenging. Longhurst (1998) identified more
than 50 distinct ecological provinces in the Ocean, defined by physical
characteristics and the structure and function of phytoplankton
communities. Longhursts scheme, however, yields far more sub-regions
than could be sensibly discussed in the space allocated within AR5.
Consequently, comparable principles were used with a division of the
non-polar ocean into seven larger sub-regions similar to Barber (1988).
It is recognized that these sub-regions do not always match physical-
chemical patterns or specific geographies, and that they interact strongly
with terrestrial regions through weather systems and the exchange of
materials. Different ocean sub-regions may also have substantially
different primary productivities and fishery catch. Notably, more than
80% of fishery catch is associated with three ocean sub-regions:
Northern Hemisphere High-Latitude Spring Bloom Systems (HLSBS),
Coastal Boundary Systems (CBS), and Eastern Boundary Upwelling
Ecosystems (EBUE; Table SM30-1, Figure 30-1). The DS (>1000 m) is
included as a separate category that overlaps with the six other ocean
sub-regions dealt with in this chapter.
30.1.2. Detection and Attribution of Climate Change and
Ocean Acidification in Ocean Sub-regions
The central goal of this chapter is to assess the recent literature on the
Ocean as a region for changes that can be attributed to climate change
and/or ocean acidification. Detailed assessments of recent physical and
chemical changes in the Ocean are outlined in WGI AR5 Chapters 2, 3, 6,
10, 13, and 14. The detection and attribution of climate change and ocean
acidification on marine organisms and ecosystems is addressed in
Chapter 6. This chapter draws on these chapters to investigate regional
changes in the physical, chemical, ecological, and socioeconomic aspects
of the Ocean and the extent to which they can be attributed to climate
change and ocean acidification.
Generally, successful attribution to climate change occurs when the full
range of possible forcing factors is considered and those related to
climate change are found to be the most probable explanation for the
detected change in question (Section 18.2.1.1). Comparing detected
changes with the expectations of well-established scientific evidence
also plays a central role in the successful attribution of detected
changes. This was attempted for seven sub-regions of the Ocean. There
are a number of general limitations to the detection and attribution of
impacts to climate change and ocean acidification that are discussed
elsewhere (Section 18.2.1) along with challenges (Section 18.2.2).
Different approaches and “best practice” guidelines are discussed in
WGI AR5 Chapters 10 and 18, as well as in several other places (Hegerl
et al., 2007, 2010; Stott et al., 2010). The fragmentary nature of ocean
observing, structural uncertainty in model simulations, the influence of
long-term variability, and confounding factors unrelated to climate
change (e.g., pollution, introduced species, over-exploitation of fisheries)
represent major challenges (Halpern et al., 2008; Hoegh-Guldberg et
al., 2011b; Parmesan et al., 2011). Different factors may also interact
synergistically or antagonistically with each other and climate change,
further challenging the process of detection and attribution (Hegerl et
al., 2007, 2010).
30.2. Major Conclusions
from Previous Assessments
An integrated assessment of the impacts of climate change and ocean
acidification on the Ocean as a region was not included in recent IPCC
assessments, although a chapter devoted to the Ocean in the Second
Assessment Report (SAR) did “attempt to assess the impacts of projected
regional and global climate changes on the oceans” (Ittekkot et al.,
1996). The fact that assessments for ocean and coastal systems are
spread throughout previous IPCC assessment reports reduces the
opportunity for synthesizing the detection and attribution of climate
change and ocean acidification across the physical, chemical, ecological,
and socioeconomic components of the Ocean and its sub-regions. The
IPCC Fourth Assessment Report (AR4) concluded, however, that, while
terrestrial sub-regions are warming faster than the oceans, “Observations
since 1961 show that the average temperature of the global ocean has
increased to depths of at least 3000 m and that the ocean has been
taking up over 80% of the heat being added to the climate system”
(AR4 Synthesis Report, p. 30). AR4 also concluded that sea levels had
risen due to the thermal expansion of the Ocean but recognized that
1663
The Ocean Chapter 30
30
our understanding of the dynamics of glaciers and ice sheets was
“too limited to assess their likelihood or provide a best estimate or an
upper boundary for sea level rise” (WGI AR4 SPM). Changes to ocean
temperature and density have been identified as having the potential
to alter large-scale ocean circulation. AR4 concluded that, with respect
to the Meridional Overturning Circulation (MOC), “it is very likely that
up to the end of the 20th century the MOC was changing significantly
at interannual to decadal time scales” (WGI AR4 Box 5.1, p. 397), despite
limited evidence of a slowing MOC.
According to AR4, Sea-level rise over the last 100 to150 years is probably
contributing to coastal erosion in many places, including the east coast
6
6
6
6
6
4
4
4
4
2
2
2
5
5
5
3
3
3
5
6
1
B
1B
1A
1A
1A
7
<1000 mDepth >1000 m
Chlorophyll a concentration (mg m
–3
)
1
A. HLSBS-
North
4. CBS
5. EBUE
3. SES
2. EUS
6. STG
0
40 80 120 160
0
5 10 15 20
1B. HLSBS-
South
2. Equatorial Upwelling Systems (EUS)
1. High-Latitude Spring Bloom Systems
(HLSBS)
1A. HLSBS-North
1B. HLSBS-South
3. Semi-Enclosed Seas (SES)
4. Coastal Boundary Systems (CBS)
5. Eastern Boundary Upwelling
Ecosystems (EBUE)
6. Subtropical Gyres (STG)
7. Deep Sea (DS; >1000 m)
Fish catch (10
6
tons yr
–1
)
(b)
(
a)
Area (10
6
km
2
)
0.01 0.03 0.1 0.3 1 3 10
Figure 30-1 | (a) Separation of the world’s oceans into seven major sub-regions (excluding an eighth area, Polar Oceans, which is considered in Chapter 28; white shaded area).
The chlorophyll-a signal measured by SeaWiFS and averaged over the period from Sep 4, 1997 to 30 Nov 2010 (NASA) provides a proxy for differences in marine productivity
(with the caveats provided in Box CC-PP). Ecosystem structure and functioning, as well as key oceanographic features, provided the basis for separating the Ocean into the
sub-regions shown. The map insert shows the distribution of Deep Sea (DS) habitat (>1000 m; Bathypelagic and Abyssopelagic habitats combined). (b) Relationship between fish
catch and area for each ocean subregion. Left panel: average fish catch (as millions tons yr
–1
) for the period 1970–2006. Right panel: surface area (millions km
2
). The top three
bars (subregions HLSBS-North, CBS, and EBUE) cover 19% of the world oceans’ area and provide 76% of the world’s fish catches. Values for fish catch, area, and primary
productivity of the ocean sub-regions are listed in Table SM30-1.
1664
Chapter 30 The Ocean
30
o
f the United States and the United Kingdom (WGII AR4 Section 1.3.3.1,
p. 92). The AR4 assessment was virtually certain that rising atmospheric
carbon dioxide (CO
2
) had changed carbonate chemistry of the ocean (i.e.,
buffering capacity, carbonate and bicarbonate concentrations), and that
a decrease in surface pH of 0.1 had occurred over the global ocean
(calculated from the uptake of anthropogenic CO
2
between 1750 and
1994; Sabine et al., 2004; Raven et al., 2005; WGI AR4 Section 5.4.2.3;
WGI AR4 Table 7.3). Large-scale changes in ocean salinity were also
observed from 1955 to 1998 and were “characterized by a global
freshening in sub-polar latitudes and salinification of shallower parts of
the tropical and subtropical oceans” (WGI AR4 Chapter 5 ES, p. 387). In
this case, freshening was observed in the Pacific, with increased salinity
being observed in the Atlantic and Indian Oceans (WGI AR4 Sections
5.3.2-5). These changes in surface salinity were qualitatively consistent
with expected changes to surface freshwater flux. Freshening of mid-
and high-latitude waters together with increased salinity at low latitudes
were seen as evidence “of changes in precipitation and evaporation
over the oceans” (WGI AR4 SPM, p. 7).
Substantial evidence presented in AR4 indicated that changing ocean
conditions have extensively influenced marine ecosystems (WGII AR4
Table 1.5). AR4 noted that there is an “accumulating body of evidence
to suggest that many marine ecosystems, including managed fisheries,
are responding to changes in regional climate caused predominately by
warming of air and sea surface temperatures (SST) and to a lesser extent
by modification of precipitation regimes and wind patterns” (WGII AR4
Section 1.3.4.2, p. 94). Observed changes in marine ecosystems and
managed fisheries reported within AR4 included changes to plankton
community structure and productivity, the phenology and biogeography
of coastal species, intertidal communities on rocky shores, kelp forests,
and the distribution of pathogens and invasive species. Changes were
also observed in coral reefs (primarily increased mass coral bleaching
and mortality) and migratory patterns and trophic interactions of marine
birds, reptiles, and mammals, as well as of a range of other marine
organisms and ecosystems (WGII AR4 Table 1.5), although a separate
exercise in detection and attribution of changes due to climate change
(as done for terrestrial studies) was not done as part of AR4.
30.3. Recent Changes and Projections
of Future Ocean Conditions
Evidence that increasing concentrations of atmospheric CO
2
have resulted
in the warming and acidification of the upper layers of the Ocean has
strengthened since AR4. Understanding the full suite of physical and
chemical changes to the Ocean is critical to the interpretation of the past
and future responses of marine organisms and ecosystems, especially
with respect to the implications for coastal and low-lying areas.
30.3.1. Physical Changes
30.3.1.1. Heat Content and Temperature
The Ocean has absorbed 93% of the extra heat arising from the
enhanced greenhouse effect (1971–2010), with most of the warming
(64%) occurring in the upper (0 to 700 m) ocean (1971–2010; WGI
A
R5 Section 3.2.3, Figure 3.2, Box 3.1). It is certain that global
average SSTs have increased since the beginning of the 20th century,
with improvements and growth of data sets and archives, and the
understanding of errors and biases since AR4 (WGI AR5 Section 2.4.2).
It is virtually certain that the upper ocean (0 to 700 m depth) has warmed
from 1971 to 2010 (Figure 30-2a), while it is likely that the surface layers
of the Ocean have warmed from the 1870s to 1971. Rates of increase
in temperature are highest near the surface of the Ocean (>0.1°C per
decade in the upper 75 m from 1971 to 2010) decreasing with depth
(0.015°C per decade at 700 m; Figure 30-2b,c). It is very likely that the
intensification of this warming near the surface has increased thermal
stratification of the upper ocean by about 4% between 0 and 200 m depth
from 1971 to 2010 in all parts of the ocean north of 40°S. It is likely
that the Ocean has warmed between 700 and 2000 m from 1957 to
2010, with the warming signal becoming less apparent or non-existent
at deeper depths (WGI AR5 Sections 3.2.1-3, Figures 3.1, 3.2, 3.9). These
changes include a significant anthropogenic signal (virtually certain;
Gleckler et al., 2012; Pierce et al., 2012), with the surface waters of all
three ocean basins warming at different rates that exceed those
expected if there were no changes to greenhouse gas (GHG) forcing
over the past century (Figure 30-2e,f,g). In this respect, the observed
record also falls within the range of historical model outputs that include
increases in the concentration of GHGs as opposed to models that do
not (Figure 30-2e,f,g).
Data archives such as Hadley Centre Interpolated SST 1.1 (HadISST1.1)
contain SSTs reconstructed from a range of sources, allowing an
opportunity to explore mean monthly, gridded, global SST from 1870
to the present (Rayner et al., 2003). The published HadISST1.1 data
set (higher temporal and spatial resolution than HadSST3) was used to
explore trends in historic SST within the sub-regions of the Ocean
(Figure 30-1a; see definition of regions in Figure SM30-1 and
Table SM30-2, column 1). The median SST for 1871–1995 from the
Comprehensive Ocean-Atmosphere Data Set (COADS) were merged
with data from the UK Met Office Marine Data Bank (MDB) to produce
monthly globally complete fields of SST on a 1° latitude-longitude SST
grid from 1870 to the present.
The surface layers of the three ocean basins have warmed (p-value
0.05, very likely), with the Indian Ocean (0.11°C per decade) warming
faster than the Atlantic (0.07°C per decade) and Pacific (0.05°C per
decade) Oceans (high confidence; Table 30-1). This is consistent with
the depth-averaged (0 to 700 m) temperature trend observed from 1971
to 2010 (Figure 30-2a).
While some regions (e.g., North Pacific) did not show a clear warming
trend, most regions showed either significant warming in the average
temperature, or significant warming in either/or the warmest and
coolest months of the year, over the period 1950–2009 (HadISST1.1
data; Table 30-1). Trends in SST show considerable sub-regional
variability (Table 30-1; Figure 30-2a). Notably, the average temperature
of most HLSBS did not increase significantly from 1950 to 2009
(except in the Indian Ocean; Table 30-1) yet the temperatures of the
warmest month (North and South Atlantic, and Southeastern Pacific)
and of the coolest month (North and South Atlantic, and South Pacific)
showed significant upward trends over this period (p-value 0.05;
Table 30-1).
1665
The Ocean Chapter 30
30
T
he two EUS warmed from 1950 to 2009 (Pacific EUS: 0.07°C per
decade, Atlantic EUS: 0.09°C per decade; Table 30-1). The average
monthly SST of the SES did not warm significantly, although the
temperature of the coolest month increased significantly within the
Baltic Sea (0.35°C per decade or 2.11°C from 1950 to 2009), as did the
temperatures of the warmest months in the Black (0.14°C per decade
o
r 0.83°C from 1950 to 2009), Mediterranean (0.11°C per decade or
0.66°C from 1950 to 2009), and Red (0.05°C per decade or 0.28°C
from 1950 to 2009) Seas over the period 1950–2009 (very likely; Table
30-1). Studies over shorter periods (e.g., 1982–2006; Belkin, 2009)
report significant increases in average SST of the Baltic (1.35°C), Black
(0.96°C), Red (0.74°C), and Mediterranean (0.71°C) Seas. Such studies
(a)
Latitude
Depth (m)
(b)
0
2
4
6
0
2
4
4
6
8
8
1
0
1
2
1
4
1
6
1
8
2
0
2
2
2
4
2
6
100
200
0
Depth (m)
(c)
1970 1980 1990 2000 2010
700
600
500
400
300
200
100
0
6.1
6.3
6.5
6.7
T0−T200 (°C)
(d)
(a,b) Temperature Trend (°C per decade)
(c) Temperature Anomaly (°C)
300
400
500
600
700
5
4
3
2
1
0
–1
–2
–0.3 –0.25 –0.2 –0.15 –0.1 –0.05 0 –0.05 –0.1 0.15 0.2 0.25 0.3
80°N60°N40°N20°N0°S20°S40°S60°S80°S
Temperature (°C)
Sea surface temperature
Historical
Natural RCP4.5
RCP8.5
OverlapOverlap
Observed
1960 1970 1980 1990 2000 2010
1960
(
e) Atlantic Ocean
(f) Indian Ocean
(g) Pacific Ocean
21002050200019501900
5
4
3
2
1
0
–1
–2
5
4
3
2
1
0
–1
–2
Temperature (°C) Temperature (°C)
21002050200019501900
21002050200019501900
Figure 30-2 | (a) Depth-averaged 0 to 700 m temperature trend for 1971–2010 (longitude vs. latitude, colors and gray contours in degrees Celsius per decade). (b) Zonally
averaged temperature trends (latitude vs. depth, colors and gray contours in degrees Celsius per decade) for 1971–2010, with zonally averaged mean temperature over plotted
(black contours in degrees Celsius). (c) Globally averaged temperature anomaly (time vs. depth, colors and gray contours in degrees Celsius) relative to the 1971–2010 mean. (d)
Globally averaged temperature difference between the Ocean surface and 200 m depth (black: annual values; red: 5-year running mean). [(a–d) from WGI AR5 Figure 3.1] (e)–(g)
Observed and simulated variations in past and projected future annual average sea surface temperature over three ocean basins (excluding regions within 300 km of the coast). The
black line shows estimates from Hadley Centre Interpolated sea surface temperature 1.1 (HadISST1.1) observational measurements. Shading denotes the 5th to 95th percentile
range of climate model simulations driven with “historical” changes in anthropogenic and natural drivers (62 simulations), historical changes in “natural drivers only (25), and the
Representative Concentration Pathways (RCPs; blue: RCP4.5; orange: RCP8.5). Data are anomalies from the 1986–2006 average of the HadISST1.1 data (for the HadISST1.1 time
series) or of the corresponding historical all-forcing simulations. Further details are given in Panels (a)-(d) originally presented in WGI AR5 Fig 3.1 and Box 21-2.
1666
Chapter 30 The Ocean
30
Continued next page
Sub-region Area
Regression slope Total change over 60 years
p-value, slope different from
zero
°C per
decade
(coolest
month)
°C per
decade
(all
months)
°C per
decade
(warmest
month)
Change
over 60
years
(coolest
month)
Change
over 60
years (all
months)
Change
over 60
years
(warmest
month)
°C per
decade
(coolest
month)
°C per
decade
(all
months)
°C per
decade
(warmest
month)
1. High-Latitude
Spring Bloom
Systems (HLSBS)
Indian Ocean 0.056 0.087 0.145 0.336 0.522 0.870 0.000 0.003 0.000
N
orth Atlantic Ocean 0.054 0.073 0.116 0.324 0.438 0.696 0.001 0.15 0.000
South Atlantic Ocean 0.087 0.063 0.097 0.522 0.378 0.582 0.000 0.098 0.000
N
orth Pacifi c Ocean (west) 0.052 0.071 0.013 0.312 0.426 0.078 0.52 0.403 0.462
North Pacifi c Ocean (east) 0.016 0.04 0.016 0.096 0.24 0.096 0.643 0.53 0.444
N
orth Pacifi c Ocean 0.033 0.055 0.015 0.198 0.33 0.09 0.284 0.456 0.319
South Pacifi c Ocean (west) 0.043 0.017 0.044 0.258 0.102 0.264 0.016 0.652 0.147
South Pacifi c Ocean (east) 0.047 0.031 0.052 0.282 0.186 0.312 0.000 0.396 0.003
S
outh Pacifi c Ocean 0.046 0.027 0.050 0.276 0.162 0.300 0.000 0.467 0.000
2. Equatorial
Upwelling Systems
(EUS)
Atlantic Equatorial Upwelling 0.101 0.090 0.079 0.606 0.540 0.474 0.000 0.000 0.000
Pacifi c Equatorial Upwelling 0.079 0.071 0.065 0.474 0.426 0.39 0.096 0.001 0.071
3. Semi-Enclosed
Seas (SES)
Arabian Gulf 0.027 0.099 0.042 0.162 0.594 0.252 0.577 0.305 0.282
Baltic Sea 0.352 0.165 0.06 2.112 0.99 0.36 0.000 0.155 0.299
Black Sea 0.004 0.053 0.139 0.024 0.318 0.834 0.943 0.683 0.009
Mediterranean Sea 0.035 0.084 0.110 0.21 0.504 0.660 0.083 0.32 0.006
Red Sea 0.033 0.07 0.047 0.198 0.42 0.282 0.203 0.138 0.042
4. Coastal
Boundary Systems
(CBS)
Atlantic Ocean (west) 0.137 0.123 0.127 0.822 0.738 0.762 0.000 0.000 0.000
Caribbean Sea/ Gulf of Mexico 0.023 0.024 0.019 0.138 0.144 0.114 0.193 0.498 0.281
Indian Ocean (west) 0.097 0.100 0.096 0.582 0.600 0.576 0.000 0.000 0.000
Indian Ocean (east) 0.099 0.092 0.080 0.594 0.552 0.480 0.000 0.000 0.000
Indian Ocean (east), Southeast
Asia , Pacifi c Ocean (west)
0.144 0.134 0.107 0.864 0.804 0.642 0.000 0.000 0.000
5. Eastern
Boundary
Upwelling
Ecosystems (EBUE)
Benguela Current 0.062 0.032 0.002 0.372 0.192 0.012 0.012 0.437 0.958
California Current 0.117 0.122 0.076 0.702 0.732 0.456 0.026 0.011 0.125
Canary Current 0.054 0.089 0.106 0.324 0.534 0.636 0.166 0.014 0.000
Humboldt Current 0.051 0.059 0.104 0.306 0.354 0.624 0.285 0.205 0.013
6. Subtropical
Gyres (STG)
Indian Ocean 0.141 0.112 0.103 0.846 0.672 0.618 0.000 0.000 0.000
North Atlantic Ocean 0.042 0.046 0.029 0.252 0.276 0.174 0.048 0.276 0.038
South Atlantic Ocean 0.079 0.083 0.098 0.474 0.498 0.588 0.000 0.017 0.000
North Pacifi c Ocean (west) 0.065 0.071 0.059 0.390 0.426 0.354 0.000 0.018 0.000
North Pacifi c Ocean (east) 0.008 0.042 0.051 0.048 0.252 0.306 0.617 0.133 0.014
North Pacifi c Ocean 0.034 0.055 0.051 0.204 0.33 0.306 0.001 0.053 0.000
South Pacifi c Ocean (west) 0.060 0.076 0.092 0.360 0.456 0.552 0.002 0.000 0.000
South Pacifi c Ocean (east) 0.055 0.056 0.088 0.330 0.336 0.528 0.000 0.058 0.000
South Pacifi c Ocean 0.056 0.060 0.089 0.336 0.360 0.534 0.000 0.027 0.000
Table 30-1 | Regional changes in sea surface temperature (SST) over the period 1950 2009 using the ocean regionalization specifi ed in Figure 30-1(a) (for further details on
regions defi ned for analysis, see Figure SM30-1 and Table SM30-2, column 1). A linear regression was fi tted to the average of all 1×1 degree monthly SST data extracted from
the Hadley Centre HadISST1.1 data set (Rayner et al., 2003) for each sub-region over the period 1950 2009. All SST values less than – 1.8°C, together with all SST pixels that
were fl agged as being sea ice, were reset to the freezing point of seawater (– 1.8°C) to refl ect the sea temperature under the ice. Separate analyses were also done to explore
trends in the temperatures extracted from the coldest-ranked and the warmest-ranked month of each year (Table SM30-2). The table includes the slope of the regression (°C per
decade), the p-value for the slope being different from zero and the total change over 60 years (i.e., the slope of linear regression multiplied by six decades) for each category.
The p-values that exceed 0.05 plus the associated slope and change values have an orange background, denoting the lower statistical confi dence in the slope being different
from zero (no slope). Note that changes with higher p-values may still describe informative trends although the level of confi dence that the slope is different from zero is lower.
1667
The Ocean Chapter 30
30
are complicated by the influence of patterns of long-term variability and
by the small size and land-locked nature of SES. Coastal Boundary Systems
(except the Caribbean and Gulf of Mexico) all showed highly significant
(p-value 0.05) warming (0.09°C to 0.13°C per decade; Table 30-1).
Among the EBUE, the Canary and Californian Current regions exhibited
a significant rate of change in the average SST (0.09°C per decade and
0.12°C per decade, respectively; p-value 0.05), while the Benguela and
Humboldt Currents did not show significant temperature changes from
1950 to 2009 (p-value ≤ 0.05; Table 30-1). There was some variability
between EUBEs in terms of the behavior of the coolest and warmest
months. The temperature of the coolest month increased significantly
from 1950 to 2009 in the case of the Benguela and California Currents
(0.06°C per decade and 0.12°C per decade, respectively; p-value 0.05),
while there was a significant increase in the temperature of the warmest
month in the case of the Canary and Humboldt Currents (0.11°C per
decade and 0.10°C per decade, respectively; Table 30-1).
The average temperature of STG showed complex patterns with increasing
temperatures (1950–2009) in the Indian, South Atlantic, and South Pacific
Oceans (very likely; 0.11°C, 0.08°C, and 0.06°C per decade, respectively;
p-value 0.05), but not in the North Atlantic or North Pacific Ocean
(p-value 0.05). These rates are half the value reported over shorter
periods (e.g., 1998–2010; Table 1 in Signorini and McClain, 2012) and
based on NOAA_OI_SST_V2 data. Given the sensitivity of coral reefs
to temperature (Eakin et al., 2010; Strong et al., 2011; Lough, 2012;
Box CC-CR), trends in key coral reef regions were also examined using
the World Resources Institute’s Reefs at Risk report (www.wri.org) to
identify HadISST1.1 grid cells containing coral reefs (Figure 30-4b).
Grouping the results into six major coral reef regions, coral reef waters
(with the notable exception of the Gulf of Mexico and Caribbean) were
found to show strong increases in average temperature (0.07°C to
0.13°C per decade) as well as the temperature of the coolest (0.07°C
to 0.14°C decade) and warmest months (very likely) (0.07°C to 0.12°C
per decade; Table 30-1). These trends in temperature have resulted in
an absolute increase in sea temperature of 0.44°C to 0.79°C from 1950
to 2009.
Given the essential role that temperature plays in the biology and
ecology of marine organisms (Box CC-MB; Sections 6.2-3; Pörtner, 2002;
Poloczanska et al., 2013), the speed of isotherm migration ultimately
determines the speed at which populations must either move, adapt,
or acclimate to changing sea temperatures (Pörtner, 2002; Burrows et
al., 2011; Hoegh-Guldberg, 2012). Burrows et al. (2011) calculated the
rate at which isotherms are migrating as the ratio of the rate of SST
change (°C yr
–1
) to the spatial gradient of temperature (°C km
–1
) over the
period 1960–2009 (Figure 30-3). Although many of these temperature
trajectories are toward the polar regions, some are not and are influenced
by features such as coastlines. This analysis and others (e.g., North
Atlantic; González-Taboada and Anadón, 2012) reveals that isotherms
in the Ocean are moving at high velocities (to over 200 km per decade),
especially at low latitudes (high confidence; Figure 30-3). Other sub-
regions showed smaller velocities with contracting isotherms (cooling)
in some areas (e.g., the Central and North Pacific, and Atlantic Oceans;
Figure 30-3). There are also changes in the timing of seasonal temperatures
in both spring and fall/autumn (Burrows et al., 2011; Poloczanska et al.,
2013), which, together with other variables (e.g., light, food availability,
geography), are likely to affect biological processes such as the
migration of species to higher latitudes, and the timing and synchrony
of reproductive and other seasonal behaviors.
Excursions of sea temperature above long-term summer temperature
maxima (or below long-term temperature minima) significantly affect
marine organisms and ecosystems (Hoegh-Guldberg, 1999; Bensoussan
et al., 2010; Crisci et al., 2011; Harley, 2011). Consequently, calculating
heat stress as a function of exposure time and size of a particular
temperature anomaly is useful in understanding recent changes to
Sub-region
Regression slope Total change over 60 years
p-value, slope different from
zero
°C per
decade
(coolest
month)
°C per
decade
(all
months)
°C per
decade
(warmest
month)
Change
over 60
years
(coolest
month)
Change
over 60
years (all
months)
Change
over 60
years
(warmest
month)
°C per
decade
(coolest
month)
°C per
decade
(all
months)
°C per
decade
(warmest
month)
Coral Reef
Provinces; see
Figure 30-4(b)
C
aribbean Sea / Gulf of Mexico 0.026 0.024 0.023 0.156 0.144 0.138 0.107 0.382 0.203
Coral Triangle and Southeast Asia 0.137 0.131 0.098 0.822 0.786 0.588 0.000 0.000 0.000
I
ndian Ocean (east) 0.081 0.097 0.116 0.486 0.582 0.696 0.000 0.000 0.000
Indian Ocean (west) 0.091 0.100 0.102 0.546 0.600 0.612 0.000 0.000 0.000
P
acifi c Ocean (east) 0.079 0.094 0.101 0.474 0.564 0.606 0.106 0.000 0.023
Pacifi c Ocean (west) 0.072 0.073 0.073 0.432 0.438 0.438 0.000 0.000 0.000
Basin Scale
North Atlantic Ocean 0.045 0.061 0.090 0.270 0.366 0.540 0.002 0.198 0.000
S
outh Atlantic Ocean 0.076 0.074 0.101 0.456 0.444 0.606 0.000 0.041 0.000
Atlantic Ocean 0.060 0.068 0.091 0.360 0.408 0.546 0.000 0.000 0.000
N
orth Pacifi c Ocean 0.030 0.052 0.046 0.180 0.312 0.276 0.000 0.248 0.006
S
outh Pacifi c Ocean 0.055 0.048 0.075 0.330 0.288 0.450 0.000 0.115 0.000
Pacifi c Ocean 0.043 0.052 0.046 0.258 0.312 0.276 0.000 0.000 0.006
Indian Ocean 0.130 0.108 0.106 0.780 0.648 0.636 0.000 0.000 0.000
Table 30-1 (continued)
1668
Chapter 30 The Ocean
30
organisms and ecosystems (e.g., coral reefs and thermal anomalies;
Strong et al., 2011). The total heat stress accumulated over the period
1981–2010 was calculated using the methodology of Donner et al.
(2007) and a reference climatology based on 1985–2000 in which the
highest monthly SST was used to define the thermal threshold, above
which accumulated thermal stress was calculated as “exposure time
multiplied by stress” or Degree Heating Months (DHM) as the running
total over 4 consecutive months. While most sub-regions of the Ocean
experienced an accumulation of heat stress (relative to a climatology
based on the period 1985–2000), equatorial and high-latitude sub-
regions in the Pacific and Atlantic Oceans have the greatest levels of
accumulated heat stress (Figure 30-4a). These are areas rich in thermally
sensitive coral reefs (Figure 30-4b; Strong et al., 2011). There was also
a higher proportion of years that have had at least one stress event
(DHM > 1) in the last 30 years (1981–2010, Figure 30-4c) than in the
preceding 30 years (1951–1980; Figure 30-4c,d).
The three ocean basins will continue warming under moderate (RCP4.5)
to high (RCP8.5) emission trajectories (high confidence) and will only
stabilize over the second half of the century in the case of low range
scenarios such as RCP2.6 (Figure 30-2e,f,g; WGI AR5 AI.4–AI.8).
Projected changes were also examined for specific ocean sub-regions using
ensemble averages from Atmosphere-Ocean General Circulation Models
(AOGCM) simulations available in the Coupled Model Intercomparison
Project Phase 5 (CMIP5) archive (Table SM30-3) for the four scenarios
of the future (RCP2.6, RCP4.5, RCP6.0, and RCP8.5; van Vuuren et al.,
2011). Ensemble averages for each RCP are based on simulations from
10 to 16 individual models (Table SM30-3). The subset of CMIP5 models
were chosen because each has historic runs enabling the derivation of
the maximum monthly mean (MMM) climatology from 1985 to 2000,
ensuring that all anomalies were comparable across time periods and
across RCPs (Figure 30-10). Model hind-cast changes matched those
observed for ocean sub-regions for the period 1980–2009 (HadISST1.1;
Figure 30-2), with the model ensemble slightly overestimating the extent
of change across the different ocean sub-regions (slope of observed/
model = 0.81, r
2
= 0.76, p-value 0.001). In this way, the absolute
amount of change projected to occur in the ocean sub-regions was
calculated for near-term (2010–2039) and long-term (2070–2099)
periods (Table SM30-4). In the near term, changes in the temperature
projected for the surface layers of the Ocean were largely indistinguishable
between the different RCP scenarios owing to the similarity in forcing
up to 2040. By the end of the century, however, SSTs across the ocean
sub-regions were 1.8°C to 3.3°C higher under RCP8.5 than those
projected to occur under RCP2.6 (Table SM30-4; Figure 30-2e,f,g). The
implications of these projected changes on the structure and function
of oceanic systems are discussed below.
30.3.1.2. Sea Level
The rate of sea level rise (SLR) since the mid-19th century has been larger
than the mean rate during the previous two millennia (high confidence).
Over the period 1901–2010, global mean sea level (GMSL) rose by 0.19
(0.17 to 0.21) m (WGI AR5 Figure SPM.3; WGI AR5 Sections 3.7, 5.6,
13.2). It is very likely that the mean rate of global averaged SLR was
1.7 (1.5 to 1.9) mm yr
–1
between 1901 and 2010, 2.0 (1.7 to 2.3) mm yr
–1
>200200
100
50
20
10 5–10 –20 –50 100 200 <–200
–5
Velocity of sea surface temperature isotherm shifts (km per decade)
Arrows indicate the direction and
magnitude of isotherm shifts
#
#
Figure 30-3 | Velocity at which sea surface temperature (SST) isotherms shifted (km per decade) over the period 1960–2009 calculated using Hadley Centre Interpolated sea
surface temperature 1.1 (HadISST1.1), with arrows indicating the direction and magnitude of shifts. Velocity of climate change is obtained by dividing the temperature trend in °C
per decade by the local spatial gradient °C km
–1
. The direction of movement of SST isotherms are denoted by the direction of the spatial gradient and the sign of the temperature
trend: toward locally cooler areas with a local warming trend or toward locally warmer areas where temperatures are cooling. Adapted from Burrows et al., 2011.
W
hite dots indicate zero or minimal
v
elocities
1669
The Ocean Chapter 30
30
b
etween 1971 and 2010, and 3.2 (2.8 to 3.6) mm yr
–1
b
etween 1993
and 2010 (WGI AR5 SPM, Section 3.7). These observations are consistent
with thermal expansion of the Ocean due to warming plus the addition
of water from loss of mass by melting glaciers and ice sheets. Current
rates of SLR vary geographically, and can be higher or lower than the
GMSL for several decades at time due to fluctuations in natural variability
and ocean circulation (Figure 30-5). For example, rates of SLR are up to
three times higher than the GMSL in the Western Pacific and Southeast
Asian region, and decreasing in many parts of the Eastern Pacific for
the period 1993–2012 as measured by satellite altimetry (Figure 30-5;
WGI AR5 Section 13.6.5).
SLR under increasing atmospheric GHG concentrations will continue for
hundreds of years, with the extent and rate of the increase in GMSL
being dependent on the emission scenario. Central to this analysis is
the millennial-scale commitment to further SLR that is likely to arise
from the loss of mass of the Greenland and Antarctic ice sheets (WGI
AR5 Section 13.5.4, Figure 13.13). SLR is very likely to increase during
t
he 21st century relative to the period 1971–2010 due to increased
ocean warming and the continued contribution of water from loss of
mass from glaciers and ice sheets. There is medium confidence that
median SLR by 2081–2100 relative to 1986–2005 will be (5 to 95%
range of process-based models): 0.44 m for RCP2.6, 0.53 m for RCP4.5,
0.55 m for RCP6.0, and 0.74 m for RCP8.5. Higher values of SLR are
possible but are not backed by sufficient evidence to enable reliable
estimates of the probability of specific outcomes. Many semi-empirical
model projections of GMSL rise are higher than process-based model
projections (up to about twice as large), but there is no consensus in
the scientific community about their reliability and there is thus low
confidence in their projections (WGI AR5 Sections 13.5.2, 13.5.3, Table
13.6, Figure 13.12).
It is considered very likely that increases in sea level will result in greater
levels of coastal flooding and more frequent extremes by 2050 (WGI
AR5 Section 13.7.2; IPCC, 2012). It is about as likely as not that the
frequency of the most intense storms will increase in some ocean basins,
(a) Total thermal stress for the period 1981–2010 (b) Coral reef provinces and locations
(c) Proportion of years with thermal stress (1951–1980) (d) Proportion of years with thermal stress (1981–2010)
Western Pacific
Eastern Pacific
Caribbean & Gulf of Mexico
Western Indian Ocean
Eastern Indian Ocean
Coral Triangle & SE Asia
0 >51 234
Average annual total of the monthly anomalies (°C)
0
Proportion of years with thermal stress (%)
100
25 50 75
The location of shallow-water coral reef cells
Figure 30-4 | Recent changes in thermal stress calculated using Hadley Centre Interpolated sea surface temperature data (HadISST1.1). A monthly climatology was created by
averaging the HadISST monthly SST values over a reference period of 1985–2000 to create 12 averages, one for each month of the year. The Maximum Monthly Mean climatology
was created by selecting the hottest month for each pixel. Anomalies were then created by subtracting this value from each sea surface temperature value, but allowing values to be
recorded only if they were greater than zero (Donner et al., 2007). Two measures of the change in thermal stress were calculated as a result: The total thermal stress for the period
1981–2010, calculated by summing all monthly thermal anomalies for each grid cell (a); and the proportion of years with thermal stress, which is defined as any year that has a
thermal anomaly, for the periods 1951–1980 (c) and 1981–2010 (d). The location of coral reef grid cells used in Table 30-1 and for comparison to regional heat stress is depicted in
(b). Each dot is positioned over a 1 × 1 degree grid cell within which lies at least one carbonate coral reef. The latitude and longitude of each reef is derived from data provided by
the World Resources Institute’s Reefs at Risk report (http://www.wri.org). The six regions are as follows: red—Western Pacific Ocean; yellow—Eastern Pacific Ocean; dark
blue—Caribbean and Gulf of Mexico; green—Western Indian Ocean; purple—Eastern Indian Ocean; and light blue—Coral Triangle and Southeast Asia.
>0
1670
Chapter 30 The Ocean
30
a
lthough there is medium agreement that the global frequency of
tropical cyclones is likely to decrease or remain constant (WGI AR5
Sections 14.6, 14.8). Although understanding of associated risks is
relatively undeveloped, coastal and low-lying areas, particularly in
southern Asia, as well as the Pacific Ocean and North Atlantic regions,
face increased flood risk (Sections 5.3.3.2, 8.2.3.3, 9.3.4.3). Future
impacts of SLR include increasing penetration of storm surges into
coastal areas and changing patterns of shoreline erosion (Section 5.3),
a
s well as the inundation of coastal aquifers by saltwater (Sections
5.4.2.5, 29.3.2). Regionally, some natural ecosystems may reduce in
extent (e.g., mangroves), although examples of habitat expansion have
been reported (Brown et al., 2011). Overall, changes to sea level are
very likely to modify coastal ecosystems such as beaches, salt marshes,
coral reefs, and mangroves (Section 5.4.2; Box CC-CR), especially where
rates of sea level rise are highest (e.g., Southeast Asia and the Western
Pacific).
−500
−250
0
250
500
Sea Level (mm)
1980
−500
−250
0
250
500
Sea Level (mm)
1960 2000 19801960 2000 19801960 2000
19801960 2000 19801960 2000 19801960 2000
−14 −12 −10
−8 −6
−4
−2
02
4
6810 12 14
Sea level change (mm yr
-1
)
San Francisco Charlottetown Stockholm
Antofagasta Manila Pago Pago
Gray lines = relative sea level changes
Red lines = global mean sea level change
Graphs Map
Figure 30-5 | Map of the rate of change in sea surface height (geocentric sea level) for the period 1993–2012 derived from satellite altimetry. Also shown are relative sea level
changes (gray lines) from selected tide gauge stations for the period 1950–2012. For comparison, an estimate of global mean sea level change is shown (red lines) with each tide
gauge time series. The relatively large short-term oscillations in local sea level (gray lines) are due to the natural climate variability and ocean circulation. For example, the large
regular deviations at Pago Pago are associated with the El Niño-Southern Oscillation. Figure originally presented in WGI AR5 FAQ 13.1, Figure 1.
1671
The Ocean Chapter 30
30
30.3.1.3. Ocean Circulation, Surface Wind, and Waves
Circulation of atmosphere and ocean (and their interactions) drives much
of the chemical, physical, and biological characteristics of the Ocean,
shaping phenomena such as ocean ventilation, coastal upwelling, primary
production, and biogeochemical cycling. Critical factors for transporting
nutrients from deep waters to the marine primary producers in the
upper layers of the ocean include wind-driven mixing and upwelling.
There has been a poleward movement of circulation features, including
a widening of the tropical belt, contraction of the northern polar vortex,
and a shift of storm tracks and jet streams to higher latitudes (medium
confidence; WGI AR5 Sections 2.7.5-6, 2.7.8; WGI AR5 Box 2.5). Long-
term patterns of variability (years to decades) continue to prevent robust
conclusions regarding long-term changes in atmospheric circulation and
winds in many cases (WGI AR5 Section 2.7.5). There is high confidence,
however, that the increase in northern mid-latitude westerly winds from
the 1950s to the 1990s, and the weakening of the Pacific Walker
Circulation from the late 19th century to the 1990s, have been largely
offset by recent changes (WGI AR5 Sections 2.7.5, 2.7.8; WGI AR5
Box 2.5). Wind stress has increased since the early 1980s over the
Southern Ocean (medium confidence; WGI AR5 Section 3.4.4), and
tropical Pacific since 1990 (medium confidence), while zonal mean wind
stress may have declined by 7% in the equatorial Pacific from 1862–
1990 due to weakening of the tropical Walker Circulation (medium
confidence; WGI AR5 Section 3.4.4; Vecchi et al., 2006). For example, it
is very likely that the subtropical gyres of the major ocean basins have
expanded and strengthened since 1993. However, the short-term nature
of observing means that these changes are as likely as not to be due to
decadal variability and/or due to longer term trends in wind forcing
associated with climate change (WGI AR5 Section 3.6). Other evidence
of changes in ocean circulation is limited to relatively short-term records
that suffer from low temporal and spatial coverage. Therefore, there is
very low confidence that multi-decadal trends in ocean circulation can
be separated from decadal variability (WGI AR5 Section 3.6.6). There is
no evidence of a long-term trend in large-scale currents such as the
Atlantic Meridional Overturning Circulation (AMOC), Indonesian
Throughflow (ITF), the Antarctic Circumpolar Current (ACC), or the
transport of water between the Atlantic Ocean and Nordic Seas (WGI
AR5 Section 3.6; WGI AR5 Figures 3.10, 3.11).
Wind speeds may have increased within the regions of EBUE (low
confidence in attribution to climate; e.g., California Current, WGI AR5
Section 2.7.2). Changing wind regimes have the potential to influence
mixed layer depth (MLD) and upwelling intensity in highly productive
sub-regions of the world’s oceans, although there is low agreement as
to whether or not upwelling will intensify or not under rapid climate
change (Bakun, 1990; Bakun et al., 2010; Box CC-UP).
Surface waves are influenced by wind stress, although understanding
trends remains a challenge due to limited data. There is medium confidence
that significant wave height (SWH) has increased since the mid-1950s
over much of the North Atlantic north of 45°N, with typical winter
season trends of up to 20 cm per decade (WGI AR5 Section 3.4.5). There
is low confidence in the current understanding of how SWH will change
over the coming decades and century for most of the Ocean. It remains
an important knowledge gap (WGI AR5 Section 3.4).
30.3.1.4. Solar Insolation and Clouds
Solar insolation plays a crucially important role in the biology of many
marine organisms, not only as a source of energy for photosynthesis
but also as a potential co-stressor in the photic zone (with temperature),
as is seen during mass coral bleaching and mortality events (e.g.,
Hoegh-Guldberg, 1999). Global surface solar insolation (from the
National Centers for Environmental Prediction–National Center for
Atmospheric Research (NCEP–NCAR) Reanalysis Project; Kalnay et al.,
1996) decreased by 4.3 W m
–2
per decade from the 1950s until 1991,
after which it increased at 3.3 W m
–2
per decade until 1999 (Ohmura,
2009; Wild, 2009), matching a broad suite of evidence from many land-
based sites (WGI AR5 Section 2.3.3). Although there is consistency
between independent data sets for particular regions, there is substantial
ambiguity and therefore low confidence in observations of global-scale
cloud variability and trends (WGI AR5 Section 2.5.6). There is also low
confidence in projections of how cloudiness, solar insolation, and
precipitation will change as the planet warms due to the large interannual
and decadal variability (El Niño-Southern Oscillation (ENSO), Pacific
Decadal Oscillation (PDO)), short observation time series, and uneven
spatial sampling, particularly in the early record (before 1950; WGI AR5
Section 2.5.8).
30.3.1.5. Storm Systems
As agents of water column mixing, storms (from small atmospheric
disturbances to intense tropical cyclones) can remix nutrients from
deeper areas into the photic zone of the Ocean, stimulating productivity.
Storms can also reduce local sea temperatures and associated stress
by remixing heat into the deeper layers of the Ocean (Carrigan and
Puotinen, 2011). Large storms can destroy coastal infrastructure and
coastal habitats such as coral reefs and mangrove forests, which can
take decades to recover (Lotze et al., 2011; De’ath et al., 2012).
Although there is low confidence for long-term trends in tropical cyclone
activity globally (largely due to the lack of reliable long-term data sets),
it is virtually certain that the frequency and intensity of the strongest
tropical cyclones in the North Atlantic have increased since the 1970s
(WGI AR5 Section 2.6.3). There is medium agreement that the frequency
of the most intense cyclones in the Atlantic has increased since 1987
(WGI AR5 Section 2.6.3) and robust evidence of inter-decadal changes in
the storm track activity within the North Pacific and North Atlantic (Lee
et al., 2012). It is also likely that there has been a decrease in the number
of land-falling tropical cyclones along the East Australian coast since
the 19th century (WGI AR5 Section 2.6.3; Callaghan and Power, 2011).
It is likely that these patterns are influenced by interannual variability
such as ENSO, with land-falling tropical cyclones being twice as common
in La Niña versus El Niño years (high confidence; Callaghan and Power,
2011). There has been an increase in the number of intense wintertime
extratropical cyclone systems since the 1950s in the North Pacific.
Similar trends have been reported for the Asian region, although analyses
are limited in terms of the spatial and temporal coverage of reliable
records (WGI AR5 Section 2.6.4). There is low confidence, however, in
large-scale trends in storminess or storminess proxies over the last century
owing to the lack of long-term data and inconsistencies between studies
(WGI AR5 Section 2.6.4).
1672
Chapter 30 The Ocean
30
30.3.1.6. Thermal Stratification
As heat has accumulated in the Ocean there has been a 4% increase in
thermal stratification of the upper layers in most ocean regions (0 to
200 m, 40-year record) north of 40°S (WGI AR5 Section 3.2.2). Increasing
thermal stratification has reduced ocean ventilation and the depth of
mixing in many ocean sub-regions (medium confidence; WGI AR5 Section
3.8.3). This in turn reduces the availability of inorganic nutrients and
consequently primary productivity (medium confidence; Section 6.3.4).
In the STG, which dominate the three major ocean basins (Section
30.5.6), satellite-derived estimates of surface chlorophyll and primary
production decreased between 1999 and 2007 (Box CC-PP). In contrast,
h
owever, in situ observations at fixed stations in the North Pacific and
North Atlantic Oceans (Hawaii Ocean Time-series (HOT) and Bermuda
Atlantic Time-series Study (BATS)) showed increases in nutrient and
chlorophyll levels and primary production over the same period, suggesting
that other processes (e.g., ENSO, PDO, North Atlantic Oscillation (NAO),
winds, eddies, advection) can counteract broad-scale trends at local
scales (Box CC-PP). The continued warming of the surface layers of the
Ocean will very likely further enhance stratification and potentially limit
the nutrient supply to the euphotic zone in some areas. The response
of upwelling to global warming is likely to vary between regions and
represents a complex interplay between local and global variables and
processes (Box CC-UW).
(a) Climatological-mean sea surface salinity (1955–2005)
(b) Annual mean evaporation–precipitation (1950–2000)
(c) The 58-year (2008 minus 1950) sea surface salinity change
(d) The 30-year (2003–2007 average centered at 2005, minus the
1960–1989 average centered at 1975) sea surface salinity difference
white areas = areas where calculations were not carried out
gray stippling = change is not significant at the 99% confidence level
white areas = areas where calculations were not carried out
gray stippling = change is not significant at the 99% confidence level
∆ Practical Salinity Scale of 1978 ∆ Practical Salinity Scale of 1978
Practical Salinity Scale of 1978 Evaporation–precipitation average (m/yr
-
1
)
32 33 34 35 36 37 38 –3 –2 –1 1 2 30
–0.25 –0.5 –0.15 –0.1 –0.05 0 0.250.20.150.10.05
–0.5 –0.4 –0.3 –2 –0.1 0 0.50.40.30.20.1
Figure 30-6 | (a) The 1955–2005 climatological-mean sea surface salinity (Antonov et al., 2010) color contoured at 0.5 Practical Salinity Scale 1978 (PSS78) intervals (black
lines). (b) Annual mean evaporation-precipitation averaged over the period 1950–2000 (National Centers for Environmental Prediction (NCEP)) color contoured at 0.5 m yr–1
intervals (black lines). (c) The 58-year (2008 minus 1950) sea surface salinity change derived from the linear trend (PSS78), with seasonal and El Niño-Southern Oscillation
(ENSO) signals removed (Durack and Wijffels, 2010) color contoured at 0.116 PSS78 intervals (black lines). (d) The 30-year (2003–2007 average centered at 2005, minus the
1960–1989 average centered at 1975) sea surface salinity difference (PSS78) (Hosoda et al., 2009) color contoured at 0.06 PSS78 intervals (black lines). Contour intervals in (c)
and (d) are chosen so that the trends can be easily compared, given the different time intervals in the two analyzes. White areas in (c) and (d) are marginal seas where the
calculations are not carried out. Regions where the change is not significant at the 99% confidence level are stippled in gray. Figure originally presented as WGI AR5 Figure 3.4.
All salinity values quoted in the chapter are expressed on the Practical Salinity Scale 1978 (PSS78) (Lewis and Fofonoff, 1979).
1673
The Ocean Chapter 30
30
30.3.2. Chemical Changes
30.3.2.1. Surface Salinity
The global water cycle is dominated by evaporation and precipitation
occurring over ocean regions, with surface ocean salinity varying with
temperature, solar radiation, cloud cover, and ocean circulation (Deser
et al., 2004). Changes in salinity influence stratification of water masses
and circulation. Ocean salinity varies regionally (Figure 30-6a) and is a
function of the balance between evaporation and precipitation (Durack
and Wijffels, 2010; WGI AR5 Section 3.3). Evaporation-dominated
regions (Figure 30-6b) such as the STG and Atlantic and Western Indian
Oceans (WGI AR5 Section 3.3.3) have elevated salinity, while areas of
high precipitation such as the North Pacific, northeastern Indian Ocean,
Southeast Asia, and the eastern Pacific have relatively low salinities
(WGI AR5 Section 3.3.3; Figure 30-6a). It is likely that large-scale trends
in salinity have also occurred in the Ocean interior, deriving from
changes to salinity at the surface and subsequent subduction (WGI AR5
Section 3.3).
Salinity trends are consistent with the amplification of the global
hydrological cycle (Durack et al., 2012; Pierce et al., 2012), a consequence
of a warmer atmosphere very likely producing the observed trend in
greater precipitation, evaporation, atmospheric moisture (Figure 30-6b),
and extreme events (WGI AR5 Sections 2.6.2.1, 3.3.4; IPCC, 2012).
Spatial patterns in salinity and evaporation-precipitation are correlated,
providing indirect evidence that these processes have been enhanced
since the 1950s (WGI AR5 Sections 3.3.2-4; WGI AR5 Figures 3.4, 3.5,
3.20d; WGI AR5 FAQ 3.3). These trends in salinity are very likely to have
a discernible contribution from anthropogenic climate change (WGI
AR5 Section 10.4.2). The combined changes in surface salinity and
temperature are consistent with changes expected due to anthropogenic
forcing of the climate system and are inconsistent with the effects of
natural climate variability, either internal to the climate system (e.g.,
ENSO, PDO; Figure 30-6c,d) or external to it (e.g., solar forcing or
volcanic eruptions; Pierce et al., 2012). There is high confidence between
climate models that the observed trends in ocean salinity will continue
as average global temperature increases (Durack and Wijffels, 2010;
Terray et al., 2012). Ramifications of these changes are largely unknown
but are of interest given the role of ocean salinity and temperature in
fundamental processes such as the AMOC.
30.3.2.2. Ocean Acidification
The Ocean has absorbed approximately 30% of atmospheric CO
2
from
human activities, resulting in decreased ocean pH and carbonate ion
concentrations, and increased bicarbonate ion concentrations (Box
CC-OA; WGI AR5 Box 3.2; WGI AR5 Figure SM30-2). The chemical response
to increased CO
2
dissolving into the Ocean from the atmosphere is
known with very high confidence (WGI AR5 Section 6.4.4). Factors such
as temperature, biological processes, and sea ice (WGI AR5 Section 6.4)
play significant roles in determining the saturation state of seawater
for polymorphs (i.e., different crystalline forms) of calcium carbonate.
Consequently, pH and the solubility of aragonite and calcite are naturally
lower at high latitudes and in upwelling areas (e.g., California Current
EBUE), where organisms and ecosystems may be relatively more exposed
t
o ocean acidification as a result (Feely et al., 2012; Gruber et al., 2012;
Figures 30-7a,b, SM30-2). Aragonite and calcite concentrations vary
with depth, with under-saturation occurring at deeper depths in the
Atlantic (calcite: 3500 to 4500 m, aragonite: 400 to 3000 m) as opposed
to the Pacific and Indian Oceans (calcite: 100 to 3000 m, aragonite: 100
to 1200 m; Feely et al., 2004, 2009; Orr et al., 2005; Figure 30-8).
Surface ocean pH has decreased by approximately 0.1 pH units since the
beginning of the Industrial Revolution (high confidence) (Figure 30-7a;
WGI AR5 Section 3.8.2; WGI AR5 Box 3.2), with pH decreasing at the
rate of –0.0013 and 0.0024 pH units yr
–1
(WGI AR5 Section 3.8.2; WGI
AR5 Table 3.2). The presence of anthropogenic CO
2
diminishes with
Arctic (>70°N)
Southern Ocean (<60°S)
Tropics (20°S–20°N)
RCP8.5
RCP2.6
(a) Surface pH
Surface pH
(b) Change in surface pH in 2090s from 1990s (RCP8.5)
1900 1950
2000
2050 2100
–0.2 –0.3 –0.4
Δ Surface pH
–0.5
7.6
7.7
7.8
7.9
8.0
8.1
8.2
Figure 30-7 | Projected ocean acidification from 11 Coupled Model Intercomparison
Project Phase 5 (CMIP5) Earth System models under RCP8.5 (other Representative
Concentration Pathway (RCP) scenarios have also been run with the CMIP5 Models):
(a) Time series of surface pH shown as the mean (solid line) and range of models
(shaded area), given as area-weighted averages over the Arctic Ocean (green), the
tropical oceans (red), and the Southern Ocean (blue). (b) Maps of the median model’s
change in surface pH from 1990s. Panel (a) also includes mean model results from
RCP2.6 (dashed lines). Over most of the Ocean, gridded data products of carbonate
system variables are used to correct each model for its present-day bias by subtracting
the model-data difference at each grid cell following (Orr et al., 2005). Where gridded
data products are unavailable (Arctic Ocean, all marginal seas and the Ocean near
Indonesia), the results are shown without bias correction. The bias correction reduces
the range of model projections by up to a factor of four; for example, in panel (a)
compare the large range of model projections for the Arctic (without bias correction) to
the smaller range in the Southern Ocean (with bias correction). Figure originally
presented in WGI AR5 Figure 6.28.
1674
Chapter 30 The Ocean
30
d
epth. The saturation horizons of both polymorphs of calcium carbonate,
however, are shoaling rapidly (1 to 2 m yr
1
, and up to 5 m yr
1
in regions
such as the California Current (Orr et al., 2005; Feely et al., 2012). Further
increases in atmospheric CO
2
are virtually certain to further acidify the
Ocean and change its carbonate chemistry (Figures SM30-2, 30-7, 30-8).
Doubling atmospheric CO
2
(~RCP4.5; Rogelj et al., 2012) will decrease
o
cean pH by another 0.1 unit and decrease carbonate ion concentrations
by approximately 100 µmol kg
1
in tropical oceans (Figure 30-8a) from
the present-day average of 250 µmol kg
1
(high confidence). Projected
changes for the open Ocean by 2100 (Figures 30-7, 30-8) range from a
pH change of –0.14 unit with RCP2.6 (421 ppm CO
2
, +1ºC, 22% reduction
of carbonate ion concentration) to a pH change of –0.43 unit with RCP8.5
1675
The Ocean Chapter 30
30
(
936 ppm CO
2
,
+3.7ºC, 56% reduction of carbonate ion concentration).
The saturation horizons will also become significantly shallower in all
oceans (with the aragonite saturation horizon between 0 and 1500 m
in the Atlantic Ocean and 0 and 600 m (poles vs. equator) in the Pacific
Ocean; Sabine et al., 2004; Orr et al., 2005; WGI AR5 Section 6.4.4; WGI
AR5 Figure 6.28). Trends toward under-saturation of aragonite and
calcite will also partly depend on ocean temperature, with surface polar
waters expected to become seasonally under-saturated with respect to
aragonite and calcite within a couple of decades (Figure 30-8c,d,e,f;
Box CC-OA; McNeil and Matear, 2008).
Overall, observations from a wide range of laboratory, mesocosm, and
field studies reveal that marine macro-organisms and ocean processes
are sensitive to the levels of ocean acidification projected under elevated
atmospheric CO
2
(high confidence; Box CC-OA, Section 6.3.2; Munday
et al., 2009; Kroeker et al., 2013). Ecosystems that are characterized by
high rates of calcium carbonate deposition (e.g., coral reefs, calcareous
plankton communities) are sensitive to decreases in the saturation states
of aragonite and calcite (high confidence). These changes are very likely
to have broad consequences such as the loss of three-dimensional coral
reef frameworks (Hoegh-Guldberg et al., 2007; Manzello et al., 2008;
Fabricius et al., 2011; Andersson and Gledhill, 2013; Dove et al., 2013)
and restructuring of food webs at relatively small (~50 ppm) additional
increases in atmospheric CO
2
. Projected shoaling of the aragonite and
calcite saturation horizons are likely to impact deep water (100 to
2000 m) communities of scleractinian corals and other benthic organisms
as atmospheric CO
2
increases (Orr et al., 2005; Guinotte et al., 2006;
WGI AR5 Section 6.4.4), although studies from the Mediterranean
and seamounts off southwest Australia report that some deep water
corals may be less sensitive (Thresher et al., 2011; Maier et al., 2013).
Organisms are also sensitive to changes in pH with respect to
physiological processes such as respiration and neural function
(Section 6.3.2). Owing to the relatively short history, yet growing effort,
to understand the implications of rapid changes in pH and ocean
carbonate chemistry, there are a growing number of organisms and
processes reported to be sensitive. The impact of ocean acidification on
marine organisms and ecosystems continues to raise serious scientific
concern, especially given that the current rate of ocean acidification (at
l
east 10 to 100 times faster than the recent glacial transitions (Caldeira
and Wickett, 2003; Hoegh-Guldberg et al., 2007)) is unprecedented
within the last 65 Ma (high confidence; Ridgwell and Schmidt, 2010)
and possibly 300 Ma of Earth history (medium confidence; Hönisch et
al., 2012; Section 6.1.2).
3
0.3.2.3. Oxygen Concentration
Dissolved O
2
is a major determinant of the distribution and abundance
of marine organisms (Section 6.3.3). Oxygen concentrations vary across
ocean basins and are lower in the eastern Pacific and Atlantic basins,
and northern Indian Ocean (Figure 30-9b; Section 6.1.1.3). In contrast,
some of the highest concentrations of O
2
are associated with cooler high-
latitude waters (Figure 30-9b). There is high agreement among analyses
providing medium confidence that O
2
concentrations have decreased
in the upper layers of the Ocean since the 1960s, particularly in the
equatorial Pacific and Atlantic Oceans (WGI AR5 Section 3.8.3; WGI AR5
Figure 3.20). A formal fingerprint analysis undertaken by Andrews et al.
(2013) concluded that recent decreases in oceanic O
2
are due to external
influences (very likely). Conversely, O
2
has increased in the North and
South Pacific, North Atlantic, and Indian Oceans, and is consistent with
greater mixing and ventilation due to strengthening wind systems (WGI
AR5 Section 3.8.3). The reduction in O
2
concentration in some areas of the
Ocean is consistent with that expected from higher ocean temperatures
and a reduction in mixing (increasing stratification) (WGI AR5 Section
3.8.3). Analysis of ocean O
2
trends over time (Helm et al., 2011b) reveals
that the decline in O
2
solubility with increased temperature is responsible
for no more than 15% of the observed change. The remaining 85%,
consequently, is associated with increased deep-sea microbial respiration
and reduced O
2
supply due to increased ocean stratification (WGI AR5
Section 6.1.1.3). In coastal areas, eutrophication can lead to increased
transport of organic carbon into adjacent ocean habitats where microbial
metabolism is stimulated, resulting in a rapid drawdown of O
2
(Weeks
et al., 2002; Rabalais et al., 2009; Bakun et al., 2010).
The development of hypoxic conditions (defined as O
2
concentrations
below ~60 µmol kg
–1
) over recent decades has been documented across
Frequently Asked Questions
FAQ 30.1 | Can we reverse the impacts of climate change on the ocean?
In less than 150 years, greenhouse gas (GHG) emissions have resulted in such major physical and chemical changes
in our oceans that it will take thousands of years to reverse them. There are a number of reasons for this. Given
its large mass and high heat capacity, the ability of the Ocean to absorb heat is 1000 times larger than that of the
atmosphere. The Ocean has absorbed at least nine-tenths of the Earths heat gain between 1971 and 2010. To reverse
that heating, the warmer upper layers of the Ocean have to mix with the colder deeper layers. That mixing can take
as much as 1000 years. This means it will take centuries to millennia for deep ocean temperatures to warm in response
to todays surface conditions, and at least as long for ocean warming to reverse after atmospheric GHG concentrations
decrease (virtually certain). But climate change-caused alteration of basic conditions in the Ocean is not just about
temperature. The Ocean becomes more acidic as more carbon dioxide (CO
2
) enters it and will take tens of thousands
of years to reverse these profound changes to the carbonate chemistry of the ocean (virtually certain). These enormous
physical and chemical changes are producing sweeping and profound changes in marine ecosystems. Large and
abrupt changes to these ecosystems are unlikely to be reversible in the short to medium term (high confidence).
1676
Chapter 30 The Ocean
30
a wide array of ocean sub-regions including some SES (e.g., Black and
Baltic Seas), the Arabian Sea, and the California, Humboldt, and
Benguela Current systems, where eruptions of hypoxic, sulfide-laden
water have also occurred in some cases (Weeks et al., 2002). Localized,
seasonal hypoxic “dead zones” have emerged in economically valuable
coastal areas such as the Gulf of Mexico (Turner et al., 2008; Rabalais
et al., 2010), the Baltic Sea (Conley et al., 2009), and the Black Sea
(Kideys, 2002; Ukrainskii and Popov, 2009) in connection with nutrient
fluxes from land. Over a vast region of the eastern Pacific stretching from
southern Chile to the Aleutian Islands, the minimum O
2
threshold (less
than 2 mg l
–1
or ~60 µmol
–1
) is found at 300 m depth and upwelling of
increasingly hypoxic waters is well documented (Karstensen et al., 2008).
Hypoxic waters in the northern Arabian Sea and Bay of Bengal are located
close to continental shelf areas. Long-term measurements reveal that
O
2
concentrations are declining in these waters, with medium evidence
that economically significant mesopelagic fish populations are being
threatened by a reduction in suitable habitat as respiratory stress
increases (Koslow et al., 2011). It should be noted that hypoxia profiles
based on a critical threshold of 60 µmol kg
–1
can convey an overly
simplistic message given that critical concentrations of O
2
in this regard
are very much species, size, temperature, and life history stage specific.
This variability in sensitivity is, however, a critical determinant for any
attempt to understand how ecosystems will respond to changing future
O
2
levels (Section 6.3.3).
There is high agreement among modeling studies that O
2
concentrations
will continue to decrease in most parts of the Ocean due to the effect
of temperature on O
2
solubility, microbial respiration rates, ocean
(a) Ocean oxygen content change (%)
(b) Oxygen concentrations in the 1990s (200–600 m)
28024080400200160120 320
1900 1950 2000 2050 2100
–4
–3
–2
–1
0
1
(mmol m
–3
)
4030–10–20–30 20100 50
–40
–50
(c) 2090s, changes from 1990s (RCP2.6)
(d) 2090s, changes from 1990s (RCP8.5)
O
2
(mmol m
–3
)
RCP2.6 RCP4.5 RCP6.0 RCP8.5Historical
Ocean oxygen content change (%)
Figure 30-9 | (a) Simulated changes in dissolved O
2
(mean and model range as shading) relative to 1990s for Representative Concentration Pathway 2.6 (RCP2.6), RCP4.5,
RCP6.0, and RCP8.5. (b) Multi-model mean dissolved O
2
(mmol m
–3
) in the main thermocline (200 to 600 m depth average) for the 1990s, and changes in the 2090s relative to
1990s for RCP2.6 (c) and RCP8.5 (d). To indicate consistency in the sign of change, regions are stippled when at least 80% of models agree on the sign of the mean change.
These diagnostics are detailed in Cocco et al. (2013) in a previous model intercomparison using the Special Report on Emission Scenarios (SRES)-A2 scenario and have been
applied to Coupled Model Intercomparison Project Phase 5 (CMIP5) models here. Models used: Community Earth System Model 1–Biogeochemical (CESM1-BGC), Geophysical
Fluid Dynamics Laboratory–Earth System Model 2G (GFDL-ESM2G), Geophysical Fluid Dynamics Laboratory–Earth System Model 2M (GFDL-ESM2M), Hadley Centre Global
Environmental Model 2–Earth System (HadGEM2-ES), Institute Pierre Simon Laplace–Coupled Model 5A–Low Resolution (IPSL-CM5A-LR), Institute Pierre Simon
Laplace–Coupled Model 5A–Medium Resolution (IPSL-CM5A-MR), Max Planck Institute–Earth System Model–Low Resolution (MPI-ESM-LR), Max Planck Institute–Earth System
Model–Medium Resolution (MPI-ESM-MR), Norwegian Earth System Model 1 (Emissions capable) (NorESM1). Figure originally presented in WGI AR5 Figure 6.30.
1677
The Ocean Chapter 30
30
v
entilation, and ocean stratification (Figure 30-9c,d; WGI AR5 Table
6.14; Andrews et al., 2013), with implications for nutrient and carbon
cycling, ocean productivity, marine habitats, and ecosystem structure
(Section 6.3.5). The outcomes of these global changes are very likely to
be influenced by regional differences in variables such as wind stress,
coastal processes, and the supply of organic matter.
30.4. Global Patterns in the Response
of Marine Organisms to Climate Change
and Ocean Acidification
Given the close relationship between organisms and ecosystems with
the physical and chemical elements of the environment, changes are
expected in the distribution and abundance of marine organisms in
response to ocean warming and acidification (Section 6.3; Boxes CC-MB,
CC-OA). Our understanding of the relationship between ocean warming
and acidification reveals that relatively small changes in temperature
and other variables can result in often large biological responses that
range from simple linear trends to more complex non-linear outcomes.
There has been an increase in studies that focus on the influence and
consequences of climate change for marine ecosystems since AR4
(Hoegh-Guldberg and Bruno, 2010; Poloczanska et al., 2013), representing
an opportunity to examine, and potentially attribute, detected changes
within the Ocean to climate change.
Evidence of global and regional responses of marine organisms to
recent climate change has been shown through assessments of multiple
studies focused on single species, populations, and ecosystems (Tasker,
2008; Thackeray et al., 2010; Przeslawski et al., 2012; Poloczanska et
al., 2013). The most comprehensive assessment, in terms of geographic
spread and number of observed responses, is that of Poloczanska et al.
(2013). This study reveals a coherent pattern in observed responses of
ocean life to recent climate change across regions and taxonomic
groups, with 81% of responses by organisms and ecosystems being
consistent with expected changes to recent climate change (high
confidence; Box CC-MB). On average, spring events in the Ocean have
advanced by 4.4 ± 0.7 days per decade (mean ± SE) and the leading
edges of species’ distributions have extended (generally poleward) by
7
2.0 ± 0.35 km per decade. Values were calculated from data series
ranging from the 1920s to 2010, although all series included data after
1990. The fastest range shifts generally occurred in regions of high
thermal velocity (the speed and direction at which isotherms move
(Burrows et al., 2011; Section 30.3.1.1)). Subsequently, Pinsky et al. (2013),
using a database of 360 fish and invertebrate species and species
groups from coastal waters around North America, showed differences
in the speed and directions that species shift can be explained by
differences in local climate velocities (Box CC-MB).
30.5. Regional Impacts, Risks, and
Vulnerabilities: Present and Future
This section explores the impacts, risks, and vulnerabilities of climate
change for the seven sub-regions within the Ocean. There is considerable
variability from region to region, especially in the extent and interaction
of climate change and non-climate change stressors. Although the latter
may complicate attribution attempts in many sub-regions, interactions
between the two groups of stressors may also represent opportunities
to reduce the overall effects on marine organisms and processes of the
environmental changes being driven by climate change (including ocean
acidification) (Crain et al., 2008; Griffith et al., 2012).
30.5.1. High-Latitude Spring Bloom Systems
High-Latitude Spring Bloom Systems (HLSBSs) stretch from 35°N to the
edge of the winter sea ice (and from 35°S to the polar front) and provide
36% of world’s fish catch (Figure 30-1b). Although much of the North
Pacific is iron limited (Martin and Fitzwater, 1988) and lacks a classical
spring bloom (McAllister et al., 1960), strong seasonal variability of
primary productivity is pronounced at all high latitudes because of
seasonally varying photoperiod and water column stability (Racault et
al., 2012). Efficient transfer of marine primary and secondary production
to higher trophic levels, including commercial fish species, is influenced
by the magnitude as well as the spatial and temporal synchrony
between successive trophic production peaks (Hjort, 1914; Cushing,
1990; Beaugrand et al., 2003; Beaugrand and Reid, 2003).
Frequently Asked Questions
FAQ 30.2 | Does slower warming in the Ocean mean less impact on plants and animals?
The greater thermal inertia of the Ocean means that temperature anomalies and extremes are lower than those
seen on land. This does not necessarily mean that impacts of ocean warming are less for the ocean than for land.
A large body of evidence reveals that small amounts of warming in the Ocean can have large effects on ocean
ecosystems. For example, relatively small increases in sea temperature (as little as 1°C to 2°C) can cause mass coral
bleaching and mortality across hundreds of square kilometers of coral reef (high confidence). Other analyses have
revealed that increased temperatures are spreading rapidly across the world’s oceans (measured as the movement
of bands of equal water temperature or isotherms). This rate of warming presents challenges to organisms and
ecosystems as they try to migrate to cooler regions as the Ocean continues to warm. Rapid environmental change
also poses steep challenges to evolutionary processes, especially where relatively long-lived organisms such as corals
and fish are concerned (high confidence).
1678
Chapter 30 The Ocean
30
30.5.1.1. Observed Changes and Potential Impacts
30.5.1.1.1. North Atlantic
The average temperature of the surface waters of the North Atlantic
HLSBS has warmed by 0.07°C per decade, resulting in an increase in
sea temperature of 0.44°C between 1950 and 2009 (likely) (p-value =
0.15; Table 30-1). Over the same period, both winter and summer
temperatures have increased significantly (0.05°C per decade and 0.12°C
per decade respectively, p-value ≤ 0.05). Since the 1970s, the Atlantic
Ocean has warmed more than any other ocean basin (0.3°C per decade;
Figure 30-2a; WGI AR5 Section 3.2.2), with greatest warming rates over
European continental shelf areas such as the southern North Sea, the
Gulf Stream front, the sub-polar gyres, and the Labrador Sea (MacKenzie
and Schiedek, 2007a,b; Levitus et al., 2009; Lee et al., 2011; González-
Taboada and Anadón, 2012). Basin-wide warming in the North Atlantic
since the mid-1990s has been driven by global warming and the current
warm phase of the Atlantic Multi-decadal Oscillation (AMO) (Wang and
Dong, 2010; WGI AR5 Section 14.7.6).
The North Atlantic is one of the most intensively fished ocean sub-
regions. The major areas for harvesting marine living resources span the
eastern North American, European, and Icelandic shelves (Livingston and
Tjelmeland, 2000). In addition, the deep regions of the Nordic Seas and
the Irminger Sea contain large populations of pelagic fish such as herring,
blue whiting, and mackerel and mesopelagic fish such as pearlsides and
redfish. The region covers a wide latitudinal range from 35°N to 80°N
and, hence, a large span in thermal habitats. This is reflected in the
latitudinal gradient from subtropical/temperate species along the
southern fringe to boreal/arctic species along the northern fringe.
Climate change is virtually certain to drive major changes to the northern
fringes of the Atlantic HLSBS by 2100. For the Barents Sea region, which
borders the HLSBS and Arctic regions, modeling projections from 1995 to
2060 (SRES B2 scenario) gave an increase in phytoplankton production
of 8%, an increase in Atlantic zooplankton production of 20%, and a
decrease of Arctic zooplankton production of 50% (Ellingsen et al., 2008).
These changes result in a total increase in zooplankton production in
the HLSBS section of the Barents Sea and a decrease in the Arctic section.
Together with poleward shifts of fish species, a substantial increase in
fish biomass and catch is also very likely at the northern fringes of the
HLSBS (Cheung et al., 2011). However, for some species such as capelin,
which feeds in summer at the ice edge and spawns in spring at the
southern Atlantic Norwegian/Murman coast of the Barents Sea, the
continuous temperature increase is very likely to cause discontinuous
changes in conditions. The limited migration potential for this small
pelagic fish is also likely to drive an eastward shift in spawning areas
to new spawning grounds along the Novaja Semlja coast (Huse and
Ellingsen, 2008).
Observations of fish and other species moving to higher latitudes (Beare
et al., 2005; Perry et al., 2005; Collie et al., 2008; Lucey and Nye, 2010)
within the North Atlantic HLSBS are consistent with results of modeling
exercises (Stenevik and Sundby, 2007; Cheung et al., 2011). Examples
from the Barents (Section 28.2.2.1), Nordic, and North Seas (Box 6-1;
Section 23.4.6) show how warming from the early 1980s influenced
North Atlantic ecosystems, where substantial biological impacts such as
l
arge-scale modification of the phenology, abundance, and distribution
of plankton assemblages and reorganization of fish assemblages have
been observed (Beaugrand et al., 2002; Edwards, 2004; Edwards and
Richardson, 2004; Tasker, 2008; Nye et al., 2009; Head and Pepin, 2010;
Simpson et al., 2011). The ranges of some cold-water zooplankton
assemblages in the northeast Atlantic have contracted towards the
Arctic since 1958, and have been replaced by warm-water zooplankton
assemblages (specifically copepods) (high confidence), which moved
up to 1000 km northward (Beaugrand et al., 2002; Beaugrand, 2009).
Although changes to surface circulation may have played a role (Reid
et al., 2001), the primary driver of the shift was shown to be regional
warming (Beaugrand et al., 2002; Beaugrand, 2004). Reorganization of
zooplankton communities and an observed decline in mean size has
implications for energy transfer to higher trophic levels including
commercial fish stocks (Beaugrand et al., 2003; Kirby and Beaugrand,
2009; Lindley et al., 2010; Section 23.4.6). Warm-water species of fish
have increased in abundance on both sides of the North Atlantic
(medium confidence; Beare et al., 2005; Collie et al., 2008; Genner et al.,
2010; Hermant et al., 2010; Lucey and Nye, 2010; Simpson et al., 2011).
The diversity of zooplankton and fish has increased as more diverse
warm-water assemblages extend northward in response to changing
environmental conditions (high confidence; Kane, 2007; Hiddink and
ter Hofstede, 2008; Beaugrand, 2009; Mountain and Kane, 2010; ter
Hofstede et al., 2010; Box 6-1; Section 23.4.6).
The past decade has been the warmest decade ever recorded in the
Barents Sea, resulting in large populations of krill shrimp and pelagic
and demersal fish stocks linked to the Atlantic and boreal ecosystem of
the Barents Sea (high confidence; Johannesen et al., 2012; Section
28.2.2.1). Recruitment to boreal fish stocks such as cod, haddock, and
herring has increased (Eriksen et al., 2012). The relatively warm Atlantic
waters have advanced northward and eastward (Årthun et al., 2012)
and sea ice has retreated along with the Arctic water masses. As a result,
boreal euphausiids, which are mainly confined to Atlantic water, have
increased in biomass and distribution (Dalpadado et al., 2012), enhancing
growth of young cod Gadus morhua (boreal) as well as the more Arctic
(arcto-boreal) capelin (Mallotus villosus). The abundance of amphipods
of more Arctic origin has decreased, resulting in poorer feeding conditions
for polar zooplankton predators such as polar cod (Boreogadus saida).
Blue whiting (Micromesistius poutassou), which spawns west of the
British Isles and feeds on zooplankton in the Norwegian Sea during the
summer, extended their summer feeding distribution into the Barents
Sea during the recent warm period.
The Norwegian Sea is one of the two core regions for the herbivore
copepod Calanus finmarchicus, an important prey species for pelagic
fish and early life stages of all fish around the rim of this high-latitude
sea including the North Sea and the Barents Sea (Sundby, 2000). C.
finmarchicus is the main food item for some of the world’s largest fish
stocks such as the Norwegian spring-spawning herring (Clupea harengus),
blue whiting (M. poutassou), and northeast Atlantic mackerel (Scomber
scombrus). These stocks have increased considerably during the recent
warming that started in the early 1980s (Huse et al., 2012). The individual
size of herring has also increased, enabling longer feeding migrations
to utilize boreal zooplankton occurring closer to distant Arctic water
masses. Mackerel (Scomber scombrus) has advanced northward and
westward into Icelandic waters (Astthorsson et al., 2012) and was even
1679
The Ocean Chapter 30
30
o
bserved in East Greenland water in summer 2013 (Nøttestad et al.,
2013). Since 2004, the sum of spawning stock biomass of the three
pelagic fish species (herring, blue whiting, and mackerel) leveled out at
around 16 million tonnes.
Observed changes in the phenology of plankton groups in the North
Sea over the past 50 years are driven by climate forcing, in particular
regional warming (high confidence; Edwards and Richardson, 2004;
Wiltshire and Manly, 2004; Wiltshire et al., 2008; Lindley et al., 2010;
Lindley and Kirby, 2010; Schluter et al., 2010), although responses are
species-specific with substantial variation within functional groups
(Edwards and Richardson, 2004; Box 6-1). For example, the peak
maximum abundance of the copepod C. finmarchicus advanced by 10
days from the 1960s to the 2000s, but its warm-water equivalent, C.
helgolandicus, did not advance (Bonnet et al., 2005). In the North Sea,
bottom temperatures in winter have warmed by 1.6°C (1980–2004;
Dulvy et al., 2008). The whole demersal fish community shifted deeper
by 3.6 m per decade over the period 1980–2004, although mean latitude
of the whole community did not show net displacement (Dulvy et al.,
2008). Within the community, cool-water specialists generally shifted
northward while abundant warm-water species shifted southward,
reflecting winter warming of the southern North Sea. The cold winter
temperatures of the shallow regions of the southern North Sea have
acted to exclude species with warm-water affinities. Trawl survey
data from the rapidly warming southern North Sea suggests waves of
immigration by southern species such as red mullet (Mullus surmuletus),
anchovy (Engraulis encrasicholus), and sardines (Sardina pilchardus),
linked to increasing population sizes and warming temperatures (Beare
et al., 2004, 2005).
In the northeast Atlantic, range expansions and contractions linked to
changing climate have also been observed in benthic crustaceans, bivalves,
gastropods, and polychaetes (medium confidence; Mieszkowska et al.,
2007; Beukema et al., 2009; Berke et al., 2010). For example, the southern
range limit of the common intertidal barnacle, Semibalanus balanoides,
contracted northward along European coastlines at a rate of 15 to 50
km per decade since 1872, and its retreat is attributed to reproductive
failure as winter temperatures warm (Southward et al., 2005; Wethey
and Woodin, 2008). Chthamalus montagui, its warm-water competitor,
increased in abundance to occupy the niche vacated by S. balanoides
(high confidence; Southward et al., 1995; Poloczanska et al., 2008).
Many of the longest and most comprehensive time series used to
investigate the ecological consequences of climate fluctuations and
fishing, that span periods of cooling and warming over the past century,
are from the northeast Atlantic (Toresen and Østvedt, 2000; Southward
et al., 2005; Sundby and Nakken, 2008; Edwards et al., 2010; Poloczanska
et al., 2013). Meta-analysis of 288 long-term data sets (spanning up to
90 years) of zooplankton, benthic invertebrates, fish, and seabirds from
the OSPAR Commission Maritime Area in the North-east Atlantic showed
widespread changes in distribution, abundance, and seasonality that
were consistent (77%) with expectations from enhanced greenhouse
warming (Tasker, 2008). The study brought together evidence of changes
in ocean climate and ecological responses across a range of species that
encompassed both exploited and unexploited species from a variety of
information types including peer-reviewed reports from International
Council for the Exploration of the Sea (ICES) Working Groups. In particular,
o
bservations indicated poleward shifts in zooplankton communities,
increasing abundance of fish species in the northern part of their ranges
and decreases in southern parts, and the expansion of benthic species
into more northerly or less coastal areas (high confidence).
The major portion of the literature on the influence of climate change
on the North Atlantic region covers time spans that are longer than for
most other sub-regions of the Ocean. Even here, however, the bulk of
the literature is limited to the last 30 to 50 years. The few publications
covering the first half of the 20th century represent an important longer
term perspective on the influence of climate change (Toresen and
Østvedt, 2000; Drinkwater, 2006; Sundby and Nakken, 2008; Bañón,
2009; Astthorsson et al., 2012). For example, distinct changes in fauna
were associated with a pronounced warming period over 1920–1940
(Wood and Overland, 2010), when fish and other fauna shifted northward
(Iversen, 1934; Southward et al., 2005; Drinkwater, 2006; Hátún et al.,
2009). The major lesson from these reports is that a rapid large-scale
temperature increase occurred in the high-latitude North Atlantic
between the 1920s and 1940s, with basin-scale consequences for
marine ecosystems that are comparable to warming and observed
impacts over the last 30 years. The former event was of great concern
within the scientific community, particularly during the late 1940s and
early 1950s (Iversen, 1934; Tåning, 1949, 1953; Southward, 1980).
However, with the subsequent long-term cooling in the 1970s, discussion
around climate responses was discontinued (Southward, 1980). The
centennial-long perspective indicates that multi-decadal variability has
played a major role in changes observed over the past 30 years. The
150-year instrumental record shows distinct warm phases of the AMO
during approximately 1930–1965 and from 1995, and cool phases
between approximately 1900–1930 and 1960–1995 (WGI AR5 Section
14.7.6). However, it is virtually certain that the enhanced warming in
recent decades cannot be explained without external forcing (WGI
AR5 Section 10.3.1.1.3). Understanding the changes in inter-decadal
variability over the next century is particularly important. The current
warm phase of the AMO is likely to terminate in the next few decades,
leading to a cooling influence in the North Atlantic and potentially
offsetting some of the effects of global warming (WGI AR5 Sections
11.3.2.4.1, 14.7.6). Over the transition period, the climate of the North
Atlantic is likely to change more rapidly than during previous transitions
since 1900.
30.5.1.1.2. North Pacific
Sub-decadal variability in the North Pacific HLSBS is dominated by ENSO
(Trenberth, 1990; WGI AR5 Section 14.4). Unlike the North Atlantic
HLSBS, the North Pacific HLSBS does not show any significant trends in
temperature over time, very likely as a consequence of climate variability
influences on long-term warming patterns (1950–2009; Table 30-1).
Decadal and longer periods of variability in the North Pacific are
reflected in the principal mode, the Pacific Decadal Oscillation (PDO;
WGI AR5 Section 14.7.3), with periodicities in SST of both 15 to 25 years
and 50 to 70 years (Minobe, 1997; Mantua and Hare, 2002). Further
modes of climate variability include the North Pacific Gyre Oscillation
(NPGO; Di Lorenzo et al., 2008; Chhak et al., 2009). The PDO exhibits
SST anomalies of one sign along the eastern boundary and the opposite
sign in western and central Pacific. The PDO has been reported to have
1680
Chapter 30 The Ocean
30
a
n anthropogenic component (Bonfils and Santer, 2011) but confidence
in this is very low (limited evidence, low agreement; WGI AR5 Section
10.3.3). The interplay of the phases of these modes of variability has strong
influence on high-latitude Pacific ecosystems (very high confidence). In the
space of 3 years, the eastern North Pacific fluctuated from one of the
warmest years in the past century (2005) to one of the coldest (2008)
(McKinnell et al., 2010; McKinnell and Dagg, 2010). This rapid change
was accompanied by large changes in primary productivity, zooplankton
communities, and fish and seabird populations (McKinnell et al., 2010;
McKinnell and Dagg, 2010; Batten and Walne, 2011; Bi et al., 2011;
Keister et al., 2011).
Climate transitions among phases of variability tend to be characterized
by abrupt reorganization of the ecosystems as dynamic trophic
relationships among species alter (Hunt et al., 2002; Peterson and
Schwing, 2003; Litzow and Ciannelli, 2007; Litzow et al., 2008; Alheit,
2009). Periods of broad-scale environmental change were observed across
high-latitude ecosystems in the North Pacific HLSBS (eastern Bering Sea
and Gulf of Alaska) during 1976–1978, 1987–1989, and 1998–1999.
These periods were associated with regime shifts in foraging fish that
occurred in 1979–1982, 1988–1992, and 1998–2001. The changes
indicate how basin-scale variability such as the PDO can manifest
across distinct ecosystems (Overland et al., 2008; Link et al., 2009a,b).
Phenological shifts observed in the zooplankton communities of the
North Pacific were very likely in response to decadal climate variability,
with distinct changes noted after the climate shifts of the 1970s and
1990s (Mackas et al., 1998; Peterson and Schwing, 2003; Chiba et al.,
2006). Modeling evidence suggests a weak shift in PDO toward more
occurrences of the negative phase but the credibility of projections
remains uncertain (WGI AR5 Section 14.7.3). It is about as likely as not
that the PDO will change its form or behavior in the future (WGI AR5
Section 14.7.3).
The Kuroshio-Oyashio Extension (KOE) in the northwest Pacific displays
pronounced decadal-scale variability (Yatsu et al., 2008; Sugisaki et
al., 2010). “Warm periods” in the mid-1970s and late 1980s were
accompanied by dramatic changes in pelagic ecosystems and sardine
and anchovy stocks (Chiba et al., 2008; Yatsu et al., 2008). Observations
and climate model simulations indicate that global warming is likely to
further alter the dynamics of the Kuroshio Current and the KOE over the
coming century (McPhaden and Zhang, 2002; Sakamoto et al., 2005;
Wu et al., 2012; Zhang et al., 2014). Alteration of the KOE will alter the
timing, magnitude, and structure of spring blooms in the western Pacific
and have implications for pelagic fish recruitment, production, and
biogeochemical cycles (Ito et al., 2004; Hashioka et al., 2009; Yatsu et
al., 2013).
Commercial catches of salmon species in the North Pacific HLSBS follow
decadal fluctuations in climate (Hare and Mantua, 2000; Mantua and
Hare, 2002). Catches peaked in the warm periods of the 1930s–1940s
and 1990s–2000s, with 2009 yielding the highest catch to date, and
warming trends are about as likely as not to have contributed to recent
peaks in some sub-regions (Morita et al., 2006; Irvine and Fukuwaka,
2011). Poleward range shifts of some large pelagic fish in the western
North Pacific, such as yellowtail Seriola quinqueradiata and Spanish
mackerel Scomberomorus niphonius, were attributed, in part, to regional
warming (high confidence) and these two species are projected to shift
3
9 to 71 km poleward from the 2000s to 2030s under SRES A1B (Tian
et al., 2012; Jung et al., 2014). Anticipating ecological responses to future
anthropogenic climate change also requires evaluation of the role of
changes to climate beyond warming per se. For example, declining sea
level pressure in the North Pacific is likely influenced by anthropogenic
forcing (Gillett et al., 2003; Gillett and Stott, 2009; WGI AR5 Section
10.3.3.4) and sea level pressure in turn is related to atmospheric climate
parameters (e.g., turbulent mixing via wind stress) that regulate
commercially significant fish populations (Wilderbuer et al., 2002).
The northern fringe of the Bering Sea is among the most productive of
marine sub-regions and includes the world’s largest single-species
fishery, walleye pollock (Theragra chalcogramma; Hunt et al., 2010).
This region underwent major changes in recent decades as a result of
climate variability, climate change, and fishing impacts (Litzow et al.,
2008; Mueter and Litzow, 2008; Jin et al., 2009; Hunt et al., 2010;
Section 28.2.2.1). Seasonal sea ice cover declined since the 1990s (to
2006), although there is no linear trend between 1953 and 2006, and
the initiation of spring ice retreat over the southeastern Bering Sea shelf
started to occur earlier (Wang et al., 2007a). Concurrent with the retreat
of the “cold pool, an area of reduced water temperature (<2°C) on
the northern Bering Sea shelf that is formed as a consequence of sea
ice and is maintained over summer (Hunt et al., 2010), bottom trawl
surveys of fish and invertebrates show a significant community-wide
northward distribution shift and a colonization of the former cold pool
areas by sub-Arctic fauna (high confidence; Wang et al., 2006a; Mueter
and Litzow, 2008).
Over a vast region of the eastern Pacific stretching from southern Chile
to the Aleutian Islands, waters low in dissolved O
2
(Oxygen Minimum
Zone (OMZ)) are found at 300 m depth (Karstensen et al., 2008).
Sporadic upwelling of these low-O
2
waters along the continental shelf
is well documented, where biological respiration can further reduce
dissolved O
2
levels and result in hypoxic or anoxic conditions that lead
to mortality of coastal fishes and invertebrates (Grantham et al., 2004;
Chan et al., 2008). The magnitude and severity of seasonal hypoxic
conditions in shallow-shelf waters of the eastern North Pacific HLSBS
increased in recent decades (Bograd et al., 2008; Chan et al., 2008). In
addition, minimum pH values in the water column usually occur near
the depths of the OMZ (WGI AR5 Box 3.2). A shoaling of the aragonite
saturation horizon has likely resulted in low-aragonite conditions within
the density layers being upwelled on the shelf of the west coast of the
USA, increasing the risk of seasonally upwelled water being relatively
acidified (Feely et al., 2008) with observed impacts on Pacific oyster
(Crassostrea gigas) hatcheries (Barton et al., 2012). In the time period
1991–2006, reductions in pH in the North Pacific between 800 and ~100
m were attributed in approximately equal measure to anthropogenic
and natural variations (Byrne et al., 2010; WGI AR5 Section 3.8.2; WGI
AR5 Figure 3.19).
30.5.1.1.3. Southern Hemisphere
The seasonal peaks in phytoplankton productivity in the Southern
Hemisphere are much less pronounced and are of smaller magnitude
than those at Northern Hemisphere high latitudes (Yoder et al., 1993).
The Southern Hemisphere HLSBS is broadly bounded by the subtropical
1681
The Ocean Chapter 30
30
f
ront and the sub-Antarctic front. Associated with the subtropical
front is intense biological activity of bloom-forming coccolithophores
(phytoplankton) (Brown and Yoder, 1994). The calcifying plankton
assemblages play a key role in carbon cycles in the region and the
transport of carbon to deep ocean sediments. The coccolithophore,
Emiliania huxleyi, extended its range south of 6in the southwest
Pacific (141°E to 145°E) over the 2 decades since 1983 (Cubillos et al.,
2007). Although the drivers for this range extension are not clear, it was
proposed that the extension is facilitated by surface warming or
changes in the abundance of grazing zooplankton.
Large regions of the sub-Antarctic surface waters are likely to become
undersaturated with respect to aragonite during winter by 2030, which
will impact calcifying plankton and Southern Ocean ecosystems (McNeil
and Matear, 2008; Bednaršek et al., 2012; Section 28.2.2.2). Shell
weights of the modern foraminifer, Globigerina bulloides, in the sediments
of the sub-Antarctic region of the HLSBS south of Australia were observed
to be 30 to 35% lower than those from sediment cores representing
preindustrial periods, consistent with a recent decline in pH (Moy et al.,
2009). Examination of the pteropod, Limacina helicina antarctica,
captured from polar waters further south shows severe levels of shell
dissolution consistent with the shoaling of the aragonite saturation
horizon and indicates that the impact of ocean acidification is already
occurring (Bednaršek et al., 2012).
While the South Pacific HLSBS has not shown warming overall, both the
warmest and coolest months show a slight, but significant, increase
over time (both 0.05°C per decade from 1950 to 2009, p-value ≤ 0.05;
Table 30-1), although some areas within this sub-region have warmed.
For example, the western Tasman Sea has shown enhanced warming
since 1900 as compared to average global trends (high confidence).
This has been driven by changes in large-scale wind-forcing leading to
a southward expansion of the South Pacific STG and intensification of
the southward-flowing East Australian Current (EAC; Cai, 2006; Hill et
al., 2008; Wu et al., 2012; WGI AR5 Section 3.6.2). Model simulations
suggest both stratospheric ozone depletion and greenhouse forcing
contribute to the observed trend in wind stress (Cai and Cowan, 2007).
Coinciding with this warming and intensified EAC is the observation
that a number of benthic invertebrates, fish, and zooplankton are now
found further south than they were in the mid-20th century (Ling, 2008;
Pitt et al., 2010; Last et al., 2011). Warming facilitated the establishment
of the grazing urchin, Centrostephanus rodgersii, in eastern Tasmania
during the late 1970s (high confidence), which has resulted in deleterious
effects on macroalgal beds (Ling, 2008; Ling et al., 2008, 2009; Banks
et al., 2010).
30.5.1.2. Key Risks and Vulnerabilities
Projected changes to the temperature of surface waters match those of
the past 50 years, with average sea temperatures in the HLSBS regions
projected to increase by 0.35°C to 1.17°C in the near term (2010–2039)
and by 1.70°C to 4.84°C over the long term (2010–2099) under the
“business as usual” (BAU) RCP8.5 scenario (Table SM30-4). Under the
lower case scenario considered here (RCP2.6), projected rates of regional
warming are much lower (0.12°C to 0.79°C) in the near term, with slight
cooling for some regions in the long term (–0.16°C to 1.46°C). Risks to
H
LSBS from warming of surface waters include changes to primary
production and carbon cycling, and the reorganization of ecosystems in
response to warmer and more acidified oceans. Both primary production
and the timing of the spring bloom in HLSBS are very sensitive to
environmental change. Latitudinal shifts in the distribution of phyto- and
zooplankton communities will alter seasonality, community composition,
and bloom dynamics (Beaugrand, 2009; Ito et al., 2010; Shoji et al., 2011).
Alteration of the structure and composition of plankton communities can
propagate through high-latitude food webs due to tight trophic linkages
(Edwards and Richardson, 2004; Beaugrand et al., 2010; Beaugrand and
Kirby, 2010). Mechanisms are complex, and tend to be non-linear, with
impacts on ecosystems, fisheries, and biogeochemical cycles being hard
to project with any certainty (Box CC-PP). A reorganization of commercial
fish stocks, with attendant social and economic disruption, is a key risk
of ongoing climate change in HLSBS sub-regions. AR4 reported that
the productivity of some marine fisheries is likely to increase in the
North Atlantic (WGII AR4 Sections 10.4.1, 12.4.7). A large number of
publications since then has substantially extended documentation of
these trends and has begun to elucidate the nuances in how marine
ecosystems and organisms respond (Sumaila et al., 2011).
An additional risk exists for sub-polar areas from the loss of seasonal
sea ice. Decreases in seasonal sea ice are very likely to lead to increases
in the length of the growth season and the intensity of the light available
to fuel phytoplankton growth and, hence, enhance primary production
and attending modifications of ecosystem structure (Arrigo et al., 2008).
In the long term, however, primary production may decrease due to the
reduced supply of nutrients to the surface layers (Box CC-PP). The decline
in Arctic sea ice will open ecological dispersal pathways, as well as new
shipping routes (Section 30.6.2.3), between the North Atlantic and the
North Pacific; large numbers of the Pacific diatom, Neodenticula seminae,
were found in the North Atlantic in 1999 (Reid et al., 2007).
HLSBSs are also vulnerable to rapid changes in the carbonate chemistry
of ocean waters. Ocean acidification will produce additional and large-
scale challenges. There is medium agreement that calcifying organisms
in these regions will be negatively affected by ocean acidification, with
substantial impacts on higher trophic levels, although there is limited
evidence at this point.
30.5.2. Equatorial Upwelling Systems
The largest upwelling systems are found in the equatorial regions of
the eastern Pacific and Atlantic Oceans (Figure 30-1a). Equatorial
Upwelling Systems (EUS) produce highly productive “cold tongues that
stretch westward across equatorial areas, which is different from other
upwelling systems (e.g., EBUE; Section 30.5.5). The associated upwelling
is a consequence of the Earth’s rotation and easterly (westward) winds
and currents, which drive water northward and southward at the northern
and southern edges of these sub-regions. As result, cold, nutrient-rich,
and high CO
2
/low pH waters are transported from the deeper layers of
the Ocean to the surface, driving high levels of primary productivity that
support 4.7% of total global fisheries productivity (Table SM30-1; Figure
30-1b). Interannual modes of variability (e.g., ENSO; WGI AR5 Section
14.4) dominate EUS, particularly in the Pacific (Barber et al., 1994;
McCarthy et al., 1996; Signorini et al., 1999; Le Borgne et al., 2002;
1682
Chapter 30 The Ocean
30
C
hristian and Murtugudde, 2003; Mestas-Nuñez and Miller, 2006;
Pennington et al., 2006; Wang et al., 2006b). Upwelling of the Pacific EUS
declines during El Niño events, when the trade winds weaken, or even
reverse, and is strengthened during La Niña events. ENSO periodicity
controls primary productivity and consequently has a strong influence
over associated fisheries production (Mestas-Nuñez and Miller, 2006).
The Intertropical Convergence Zone (ITCZ; WGI AR5 Section 14.3.1.1),
an important determinant of regional ocean temperature, is located at
the edges of the Indian and Pacific equatorial upwelling zone and
influences a range of variables including productivity, fisheries, and
precipitation. The EUS are also affected by inter-decadal variability (e.g.,
Inter-decadal Pacific Oscillation (IPO); Power et al., 1999; WGI AR5
Section 14.3).
30.5.2.1. Observed Changes and Potential Impacts
The average sea temperature associated with the EUS has increased
significantly (p-value 0.05), by 0.43°C and 0.54°C from 1950 to 2009
in the Pacific and Atlantic EUS, respectively (Table 30-1). In the Pacific,
regional variability in SST trends is driven by the temporal patterns in
ENSO and the more frequent El Niño Modoki or Central Pacific El Niño
events in recent decades (high confidence; Ashok et al., 2007; Yu and
Kao, 2007; Lee and McPhaden, 2010; WGI AR5 Section 14.2.4.4). The
faster warming of the Atlantic EUS is likely to be associated with a
weakening of upwelling (Tokinaga and Xie, 2011). SLR in the eastern
equatorial Pacific has been decreasing by up to −10 mm yr
–1
since 1993
(Church et al., 2006; Figure 30-5).
Coral reefs in the EUS of the eastern Pacific (e.g., Galápagos and Cocos
Islands) have relatively low species diversity and poorly developed
carbonate reef frameworks, due to the low pH and aragonite saturation
of upwelling waters (high confidence; Glynn, 2001; Manzello et al.,
2008; Manzello, 2010). Prolonged periods of elevated temperature
associated with El Niño have negatively affected corals, kelps, and
associated organisms, and resulted in several possible local extinctions
(high confidence; Glynn, 2011). Since 1985, coral reefs from west of
South America to the Gilbert Islands of Kiribati have experienced the
h
ighest levels of thermal stress relative to other areas (Donner et al.,
2010). In 1982/1983, mass coral bleaching and mortality affected most
of the reef systems within the eastern equatorial Pacific (Glynn, 1984;
Baker et al., 2008). Subsequent canonical El Niño and Central Pacific
El Niño events in 1997/1998, 2002/2003, 2004/2005, and 2009/2010
(WGI AR5 Section 14.4.2; WGI AR5 Figure 14.13) triggered mass coral
bleaching by adding to the background increases in sea temperatures
(high confidence; Donner et al., 2010; Obura and Mangubhai, 2011;
Vargas-Ángel et al., 2011). In some locations, impacts of El Niño have also
interacted with other anthropogenic changes, such as those arising from
changes to fishing pressure (Edgar et al., 2010), further complicating
the attribution of recent ecological changes to climate change.
30.5.2.2. Key Risks and Vulnerabilities
Climate models indicate that ENSO is virtually certain to continue to be
a major driver of oceanic variability over the coming century, although
not all models can accurately replicate its behavior (WGI AR5 Section
9.5.3). Superposition of a warming ocean on future ENSO activity
(possibly modified in frequency and intensity) is likely to result in oceanic
conditions that are different from those experienced during past El Niño
and La Niña events (Power and Smith, 2007). Temperatures within EUS
sub-regions are projected to continue to warm significantly (p-value ≤
0.05). Under RCP8.5, SST of the Atlantic EUS is projected to increase by
0.81°C over 2010–2039 and 2.56°C over 2010–2099, with similar
increases projected for the Pacific EUS (Table SM30-4). Differences
between RCPs for the two EUS become clear beyond mid-century, with
warming of SST over 2010–2099 being 0.43°C and 0.46°C under
RCP2.6, and 3.01°C and 3.03°C under RCP8.5, for Pacific and Atlantic
EUS respectively (Table SM30-4). These projected increases in sea
temperature will increase heat stress and ultimately irreversibly degrade
marine ecosystems such as coral reefs (very likely). Further increases in
atmospheric CO
2
will cause additional decrease in pH and aragonite
saturation of surface waters (adding to the low pH and aragonite
saturation of upwelling conditions), with significant differences between
emission trajectories by the middle of the century. These changes in
ocean carbonate chemistry are very likely to negatively affect some
Frequently Asked Questions
FAQ 30.3 | How will marine primary productivity change with ocean warming and acidification?
Drifting microscopic plants known as phytoplankton are the dominant marine primary producers at the base of
the marine food chain. Their photosynthetic activity is critically important to life in general. It provides oxygen,
supports marine food webs, and influences global biogeochemical cycles. Changes in marine primary productivity
in response to climate change remain the single biggest uncertainty in predicting the magnitude and direction of
future changes in fisheries and marine ecosystems (low confidence). Changes have been reported to a range of
different ocean systems (e.g., High-Latitude Spring Bloom Systems, Subtropical Gyre Systems, Equatorial Upwelling
Systems, and Eastern Boundary Upwelling Ecosystems), some of which are consistent with changes in ocean
temperature, mixing, and circulation. However, direct attribution of these changes to climate change is made difficult
by long-term patterns of variability that influence productivity of different parts of the Ocean (e.g., Pacific Decadal
Oscillation). Given the importance of this question for ocean ecosystems and fisheries, longer time series studies
for understanding how these systems are changing as a result of climate change are a priority (high agreement).
1683
The Ocean Chapter 30
30
m
arine calcifiers, although many of the species from this region are
adapted to the low aragonite and calcite saturation states that result
from equatorial upwelling, albeit with much lower rates of calcification
(Manzello, 2010; Friedrich et al., 2012). A substantial risk exists with
respect to the synergistic interactions between sea temperature and
declining pH, especially as to how they influence a large number of key
biological processes (Box CC-OA).
There is low confidence in the current understanding of how (or if)
climate change will influence the behavior of ENSO and other long-term
climate patterns (Collins et al., 2010; WGI AR5 Section 12.4.4.2). There
is also low agreement between different CMIP5 General Circulation
Models (GCMs) on how ocean warming will affect ENSO, with no
significant change to ENSO amplitude in half of the models examined,
and both increasing and decreasing activity in others (Guilyardi et al.,
2012). These differences appear to be a consequence of the delicate
balance within ENSO between dampening and amplifying feedbacks,
and the different emphasis given to these processes within the different
GCMs (Collins et al., 2010). Other studies have looked at the interaction
between the STG and EUS, and warming of surface waters in the Pacific,
with at least one study projecting the possible expansion of the STG at
the expense of the EUS (Polovina et al., 2011). In the latter case, the
area of equatorial upwelling within the North Pacific would decrease
by 28%, and primary production and fish catch by 15%, by 2100. Many
of the projected changes imply additional consequences for pelagic
fisheries resulting from the migration of fish stocks deriving from
changing distribution of particular sea temperatures (Lehodey et al., 2006,
2008, 2011; Cheung et al., 2010; Sumaila et al., 2011; Bell et al., 2013b).
These projections suggest that fisheries within EUS will experience
increased vulnerability as a result of climate change (low confidence).
30.5.3. Semi-Enclosed Seas
Semi-Enclosed Seas (SES) represent a subset of ocean sub-regions
that are largely land locked and consequently heavily influenced by
surrounding landscapes and local climates (Healy and Harada, 1991). In
most cases, they support small but regionally significant fisheries (3.3%
of global production; Table SM30-1; Figure 30-1b) and opportunities for
other industries such as tourism. Five SES (all over 200,000 km² with
single entrances <120 km wide) are considered here. This particular
geography results in reduced circulation and exchange with ocean waters,
and jurisdictions for these water bodies that are shared by two or more
neighboring states. In many cases, the small volume and disconnected
nature of SES (relative to coastal and oceanic environments) makes
them highly vulnerable to both local and global stressors, especially with
respect to the much reduced options for the migration of organisms as
conditions change.
30.5.3.1. Observed Changes and Potential Impacts
30.5.3.1.1. Arabian Gulf
The Arabian Gulf (also referred to as the Persian Gulf), along with the
Red Sea, is the world’s warmest sea, with both extreme negative and
positive temperature excursions (annual temperature range of 12°C to
3
5°C). Like other SES, the Arabian Gulf is particularly vulnerable to
changing environmental conditions as a result of its landlocked nature.
Trends in SST were not significant over the period 1950–2009 (Table
30-1), which is probably due to long-term variability, and a consequence
of regional and abrupt changes that occurred in the late 1980s (Conversi
et al., 2010). In keeping with this, recent (1985–2002) localized analyses
(e.g., Kuwait Bay) show strong and significant warming trends (based
in this case on Advanced Very High Resolution Radiometer (AVHRR)
National Oceanic and Atmospheric Administration (NOAA) satellite data)
of 0.6°C per decade (Al-Rashidi et al., 2009). There is limited evidence
and low agreement as to how this variability influences marine ecosystems
and human activities within the Arabian Gulf, although impacts on some
ecosystem components (e.g., coral reefs) have been defined to some
extent. The mass coral bleaching and mortality that occurred in 1996
and 1998 were a direct result of the sensitivity of reef-building corals
to unusually elevated sea temperatures (high confidence; Riegl, 2002,
2003; Box CC-CR). These changes to coral reefs have resulted in a loss
of fish species that feed on coral-associated invertebrates while
herbivores and planktivorous fish abundances have increased (medium
confidence; Riegl, 2002). Despite coral ecosystems in this sub-region
being adapted to some of the highest temperatures in shallow seas on
Earth, anthropogenic climate change is driving higher frequencies and
intensities of mass coral bleaching and mortality (Riegl et al., 2011).
Other biological changes (e.g., harmful algal blooms and fish kills; Heil
et al., 2001) have been associated with the increasing sea temperatures
of the Arabian Gulf, although attribution to increasing temperatures as
opposed to other factors (e.g., water quality) is limited (Bauman et al.,
2010).
30.5.3.1.2. Red Sea
Few studies have focused on attributing recent changes in Red Sea
ecosystems to climate change (including ocean acidification). The Red
Sea warmed by 0.74°C from 1982 to 2006 (Belkin, 2009), although
trends in the average SST, however, are not significant from 1950 to
2009 (p-value > 0.05; Table 30-1) owing to a high degree of variability
involved when longer periods were examined (supplementary material
in Belkin, 2009). The temperature of the warmest month of the year,
however, showed a significant increase over the 60-year period (0.05°C
per decade; Table 30-1). Regional trends within the Red Sea may also
differ, with at least one other study reporting higher rates of warming
for the central Red Sea (1.46°C, relative to 1950–1997 NOAA Extended
Reconstructed SST (ERSST) v3b climatology; Cantin et al., 2010).
Long-term monitoring of coral community structure and size over 20
years shows that average colony size of corals has declined (high
confidence) and species’ latitudinal limits may have changed (medium
confidence). The decline in average colony size is ascribed to heat-
mediated bleaching as well as increases in coral diseases and crown of
thorns starfish (Acanthaster sp.) predation (Riegl et al., 2012). The
patterns of this decline correlate well with the pattern of recent heating
in the Red Sea (Raitsos et al., 2011), with the biggest changes being
seen in the southern part of the Red Sea. Skeletal growth of the long-
lived massive coral Diploastrea heliopora has declined significantly, very
likely due to warming temperatures (medium confidence; p-value 0.05;
Cantin et al., 2010).
1684
Chapter 30 The Ocean
30
C
antin et al. (2010) proposed that the massive coral Diploastrea helipora
will cease to grow in the central Red Sea by 2070 under SRES A1B and
A2 (medium confidence), although this may not hold for other coral
species. For example, an increase in linear extension of Porites corals,
beginning in the 1980s, was recorded in the northern Red Sea (Heiss,
1996), where temperatures have increased by 0.74ºC from 1982 to 2006
(Belkin, 2009), suggesting that these corals were living in sub-optimal
conditions (cooler waters). They may therefore benefit from elevated
temperature before reaching their thermal threshold, at which point
growth rates would be predicted to decline, as they are doing in other
oceans. Riegl and Piller (2003) concluded that coral habitats at moderate
depths in the Red Sea might provide important refugia from some aspects
of climate change in the future (limited evidence). Silverman et al.
(2007) quantified the sensitivity of net coral reef ecosystem calcification
to changes in carbonate chemistry (pH, aragonite saturation). Their
results demonstrate a strong negative effect of ocean acidification on
ecosystem-scale calcification and decalcification, and show that small
changes in carbonate dissolution could have large-scale implications
for the long-term persistence of carbonate coral reef systems within the
Red Sea (Silverman et al., 2007, 2009).
30.5.3.1.3. Black Sea
The temperature of the surface waters of the Black Sea increased by
0.96ºC from 1982 to 2006 (Belkin, 2009), which is consistent with other
studies (high confidence; Buongiorno Nardelli et al., 2010; Bozkurt and
Sen, 2011). As with other SES (i.e., Arabian Gulf and Baltic, Mediterranean,
and Red Seas), longer data sets do not reveal a significant trend due to
large-scale variability prior to 1982, which may be due to the influence
of AMO, NAO, and other long-term sources of variability (Table 30-1;
supplementary material in Belkin, 2009). Buongiorno Nardelli et al.
(2010) observed that short-term SST variability (week-month) is strongly
influenced by interactions with the overlying atmosphere, which itself
is strongly influenced by the surrounding land temperatures. As with
the Mediterranean and Red Seas, however, a significant upward trend
in the temperature is recorded in the warmest month of the year over
the period 1950–2009 (Table 30-1). Freshwater discharge from rivers
draining into the Black Sea has remained more or less constant since
the early 1960s (Ludwig et al., 2009). Increasing water temperature has
steadily eliminated the Cold Intermediate Layer (CIL; temperatures below
8°C) throughout the Black Sea basin over 1991–2003 (high confidence;
Oguz et al., 2003). Reduced water column mixing and upwelling during
warmer winter periods has reduced the supply of nutrients to the upper
layers of the Black Sea (Oguz et al., 2003) and expanded areas of low
O
2
in the deeper parts of the Black Sea, which is the world’s largest
anoxic marine basin (high confidence; Murray et al., 1989). These
changes coincided with the collapse of fish stocks and the invasion by
the ctenophore, Mnemiopsis leidyi, in the 1980s (Oguz et al., 2008),
while inputs of nutrients such as phosphate from the Danube River
has decreased strongly since 1992–1993 (Oguz and Velikova, 2010).
Environmental perturbations explain the declining levels of primary
productivity, phytoplankton, bacterioplankton, and fish stocks in the
Black Sea from the mid-1990s (Yunev et al., 2007; Oguz and Velikova,
2010). The Black Sea system is very dynamic and is strongly affected by
non-climate stressors in addition to climate change, making attribution
of detected trends to climate change difficult.
30.5.3.1.4. Baltic Sea
Temperatures in the highly dynamic Baltic Sea increased substantially
since the early 1980s (Aleksandrov et al., 2009; Belkin, 2009), with
increases of 1.35°C (1982–2006) being among the highest rate of change
seen in any SES (Belkin, 2009). Increases of this magnitude are not seen
in longer records throughout the Baltic Sea (1861–2001: MacKenzie et
al., 2007; MacKenzie and Schiedek, 2007a,b; 1900–1998: Madsen and
Højerslev, 2009). The salinity of the surface and near bottom waters of
the Baltic Sea, for example, Gdansk Basin (Aleksandrov et al., 2009)
and central Baltic (Fonselius and Valderrama, 2003; llmann et al.,
2003), decreased from 1975 to 2000, due to changing rainfall and river
runoff, and a reduction in the pulses of seawater (vital for oxygenation
and related chemical changes) from the North Sea through its opening
via the Kattegat (high confidence; Samuelsson, 1996; Conley et al., 2009;
Hänninen and Vuorinen, 2011). There is a strong vertical zonation within
the Baltic Sea in terms of the availability of O
2
. The shallow sub-regions
of the Baltic are relatively well oxygenated. However, O
2
levels are low
in the deeper basins, producing conditions in which organisms and
ecosystems are exposed to prolonged hypoxia.
The annual biomass of phytoplankton has declined almost threefold in
the Baltic Transition Zone (Kattegat, Belt Sea) and Western Baltic Sea
since 1979 (Henriksen, 2009), reputedly due to changing nitrogen loads
in the Danish Straits (medium confidence) in addition to increasing
sea temperature (very likely; Madsen and Højerslev, 2009). Reduced
phytoplankton production may have decreased the productivity of
fisheries in the western Baltic Sea and the Transition Zone (low to
medium confidence; Chassot et al., 2007). Decreasing salinity in the Baltic
deep basins may also affect zooplankton reproduction, especially that of
the copepod Pseudocalanus acuspes, contributing to density-dependent
decrease in growth of the commercially important herring and sprat
stocks (high confidence; Möllmann et al., 2003, 2005; Casini et al., 2011).
The strong relationship between phytoplankton and fish production, and
increasing sea temperature, decreasing salinity, and other environmental
factors, suggests that major changes in fisheries production will occur
as sea temperatures increase and the hydrological cycle in the Baltic
region changes (high confidence; MacKenzie et al., 2012). A combination
of climate change-induced oceanographic changes (i.e., decreased salinity
and increased temperatures), eutrophication, and overfishing have
resulted in major changes in trophic structure in the deep basins of the
Baltic Sea (Möllmann et al., 2009). This had important implications
for cod, a commercially important top predator (medium confidence;
Lindegren et al., 2010).
30.5.3.1.5. Mediterranean Sea
The Mediterranean Sea is strongly linked to the climates of North Africa
and Central Europe. SST within the Mediterranean increased by 0.43°C
from 1957 to 2008 (supplementary material in Belkin, 2009), although
analysis of data from 1950 to 2009 detected only a significant trend in
summer temperature (0.11°C per decade, p-value 0.05; Table 30-1)
due to large fluctuations in SST prior to the 1980s. Surface temperatures
increased in the Mediterranean Sea consistent with significant increases
in SST at a number of monitoring sites (robust evidence, high agreement;
e.g., Coma et al., 2009; Conversi et al., 2010; Calvo et al., 2011). It is
1685
The Ocean Chapter 30
30
l
ikely that temperatures, along with salinity, have also increased at
depth (400 m or more) in the western Mediterranean Sea over the past
30 to 40 years which, when analyzed in the context of heat budget and
water flux of the Mediterranean, is consistent with anthropogenic
greenhouse warming (Bethoux et al., 1990; Rixen et al., 2005; Vargas-
Yáñez et al., 2010). Large-scale variability such as the AMO and NAO
can obscure or accentuate the overall warming trend (Marullo et al.,
2011; WGI AR5 Sections 14.5.1, 14.7.6). Relatively warm episodes in the
1870s, 1930–1970s, and since the mid-1990s, for example, exhibit an
influence of the AMO (Kerr, 2000; Moron, 2003). Reported temperature
anomalies in the Mediterranean, often locally manifesting themselves
as periods of low wind, increased water column stratification, and a
deepening thermocline, are associated with positive phases of the NAO
index (Molinero et al., 2005; Lejeusne et al., 2010).
Sea levels have increased rapidly in some areas over recent decades
and are also strongly influenced by NAO phases. The rate has been
approximately 3.4 mm yr
–1
(1990–2009) in the northwest Mediterranean
(high confidence; Calvo et al., 2011). These influences are reduced when
measurements are pooled over longer time scales, resulting in a lower
rate of SLR (Massuti et al., 2008). If the positive phase of the NAO is
more frequent in the future (Terray et al., 2004; Kuzmina et al., 2005;
WGI AR5 Section 14.4.2), then future SLR may be slightly suppressed
as a result of atmospheric influences (medium confidence; Jordà et al.,
2012). As temperatures have increased, the Mediterranean has become
more saline (+0.035 to 0.040 psu from 1950 to 2000; Rixen et al., 2005)
with the length of the thermal stratification period persisting twice as
long in 2006 as it did in 1974 (Coma et al., 2009).
Conditions within the Mediterranean Sea changed abruptly and
synchronously with similar changes across the North, Baltic, and Black
Seas in the late 1980s (Conversi et al., 2010), which possibly explains the
lack of trend in SES SST when examined from 1950 to 2009 (Table 30-1).
These changes in physical conditions (increased temperature, higher sea
level pressure, positive NAO index) also coincided with step changes in
the diversity and abundance of zooplankton, decreases in stock abundance
of anchovies and the frequency of “red tides, and increases in mucilage
outbreaks (Conversi et al., 2010). Mucilage outbreaks are strongly
associated with warmer and more stratified water columns (high
confidence), and lead to a greater abundance and diversity of marine
microbes and potentially disease-causing organisms (likely; Danovaro
et al., 2009). Increasing temperatures are also driving the northward
spread of warm-water species (medium confidence) such as the sardine
Sardinella aurita (Sabatés et al., 2006; Tsikliras, 2008), and have
contributed to the spread of the invading Atlantic coral Oculina patagonia
(Serrano et al., 2013). The recent spread of warm-water species that have
invaded through the Straits of Gibraltar and the Suez Canal into cooler
northern areas is leading to the “tropicalization of Mediterranean
fauna (high confidence; Bianchi, 2007; Ben Rais Lasram and Mouillot,
2008; CIESM, 2008; Galil, 2008, 2011). Warming since the end of the
1990s has accelerated the spread of tropical invasive species from the
eastern Mediterranean basin (Raitsos et al., 2010; Section 23.6.5).
In addition to general warming patterns, periods of extreme temperatures
have had large-scale and negative consequences for Mediterranean
marine ecosystems. Unprecedented mass mortality events, which affected
at least 25 prominent invertebrate species, occurred during the summers
o
f 1999, 2003, and 2006 across hundreds of kilometers of coastline in
the northwest Mediterranean Sea (very high confidence; Cerrano et al.,
2000; Garrabou et al., 2009; Calvo et al., 2011; Crisci et al., 2011). Events
coincided with either short periods (2 to 5 days: 2003, 2006) of high
sea temperatures (27°C) or longer periods (30 to 40 days) of modestly
high temperatures (24°C: 1999; Bensoussan et al., 2010; Crisci et al.,
2011). Impacts on marine organisms have been reported in response
to the extreme conditions during these events (e.g., gorgonian coral
mortality; Coma et al., 2009), shoot mortality, and anomalous flowering
of seagrasses (high confidence; Diaz-Almela et al., 2007; Marbà and
Duarte, 2010). The frequency and intensity of these types of heat
stress events are expected to increase as sea temperatures increase
(high confidence).
Longer-term data series (over several decades) of changes in relative
acidity of the Mediterranean Sea are scarce (Calvo et al., 2011; The
MerMex Group, 2011). Recent re-analysis, however, has concluded that
the pH of Mediterranean waters has decreased by 0.05 to 0.14 pH units
since the preindustrial period (medium confidence; Luchetta et al., 2010;
Touratier and Goyet, 2011). Anthropogenic CO
2
has penetrated the
entire Mediterranean water column, with the western basin being more
contaminated than the eastern basin (Touratier and Goyet, 2011).
Studies that have explored the consequences of ocean acidification for
the biology and ecology of the Mediterranean Sea are rare (Martin and
Gattuso, 2009; Rodolfo-Metalpa et al., 2010; Movilla et al., 2012),
although insights have been gained by studying natural CO
2
seeps at
Mediterranean sites such as Ischia in Italy, where biodiversity decreases
with decreasing pH toward the vents, with a notable decline in calcifiers
(Hall-Spencer et al., 2008). Transplants of corals, molluscs, and bryozoans
along the acidification gradients around seeps reveal a low level of
vulnerability to CO
2
levels expected over the next 100 years (low
confidence; Rodolfo-Metalpa et al., 2010, 2011). However, periods of high
temperature can increase vulnerability to ocean acidification, thereby
increasing the long-term risk posed to Mediterranean organisms and
ecosystems as temperatures warm. Significantly, some organisms such
as seagrasses and some macroalgae appeared to benefit from local
ocean acidification (Hall-Spencer et al., 2008).
30.5.3.2. Key Risks and Vulnerabilities
SES are highly vulnerable to changes in global temperature on account
of their small volume and landlocked nature. Consequently, SES will
respond faster than most other parts of the Ocean (high confidence).
Risks to ecosystems within SES are likely to increase as water columns
become further stratified under increased warming, promoting hypoxia
at depth and reducing nutrient supply to the upper water column
(medium evidence, high agreement). The impact of rising temperatures
on SES is exacerbated by their vulnerability to other human influences
such as over-exploitation, pollution, and enhanced runoff from modified
coastlines. Due to a mixture of global and local human stressors, key
fisheries have undergone fundamental changes in their abundance and
distribution over the past 50 years (medium confidence). A major risk
exists for SES from projected increases in the frequency of temperature
extremes that drive mass mortality events, increasing water column
stabilization leading to reduced mixing, and changes to the distribution
and abundance of marine organisms. The vulnerability of marine
1686
Chapter 30 The Ocean
30
e
cosystems, fisheries, and human communities associated with the SES
will continue to increase as global temperatures increase.
Sea temperatures are very likely to increase in the five SES under
moderate (RCP6.0) to high (RCP8.5) future scenarios. Under BAU (RCP8.5;
Table SM30-3), sea temperatures in the SES are projected to increase by
0.93°C to 1.24°C over 2010–2039 (Table SM30-4). Increases of 3.45°C
to 4.37°C are projected over 2010–2099, with the greatest increases
projected for the surface waters of the Baltic Sea (4.37°C) and Arabian
Gulf (4.26°C), and lower yet substantial amounts of warming in the Red
Sea (3.45°C) (Table SM30-4). The heat content added to these small
ocean regions is very likely to increase stratification, which will reduce
the nutrient supply to the upper layers of the water column, reducing
primary productivity and driving major changes to the structure and
productivity of fisheries. Reduced mixing and ventilation, along with
increased microbial metabolism, will very likely increase hypoxia and
expand the number and extent of dead zones.” Changing rainfall
intensity (Section 23.3; WGI AR5 Section 12.4.5) can exert a strong
influence on the physical and chemical conditions within SES, and in
some cases will combine with other climatic changes to transform these
areas. These changes are likely to increase the risk of reduced bottom-
water O
2
levels to Baltic and Black Sea ecosystems (due to reduced
solubility, increased stratification, and microbial respiration), which is
very likely to affect fisheries. These changes will increase the frequency
and intensity of impacts arising from heat stress, based on responses
to temperature extremes seen over the past 30 years, such as the mass
mortality of benthic organisms that occurred in the Mediterranean Sea
during the summers of 1999, 2003, and 2006, and the Arabian Gulf in
1996 and 1998. Extreme temperature events such as heat waves are
projected to increase (high confidence; Section 23.2; IPCC, 2012).
Projections similar to those outlined in Section 30.5.4.2 can be applied to
the coral reefs of the Arabian Gulf and the Red Sea, where temperatures
are very likely to increase above established thresholds for mass coral
bleaching and mortality (very high confidence; Figure 30-10).
30.5.4. Coastal Boundary Systems
The Coastal Boundary Systems (CBS) are highly productive regions,
comprising 10.6% of primary production and 28.0% of global fisheries
production (Table SM30-1; Figure 30-1b). The CBS include the marginal
seas of the northwest Pacific, Indian, and Atlantic Oceans, encompassing
the Bohai/Yellow Sea, East China Sea, South China Sea, and Southeast
Asian Seas (e.g., the Timor, Arafura, and Sulu Seas, and the northern
coast of Australia) in the Pacific; the Arabian Sea, Somali Current system,
East Africa coast, Mozambique Channel, and Madagascar in the Indian
Ocean; and the Caribbean Sea and Gulf of Mexico in the Atlantic Ocean).
Some CBS are dominated by powerful currents such as the Kuroshio
(Pacific), or are strongly influenced by monsoons (e.g., Asian-Australian
and African monsoons).
30.5.4.1. Observed Changes and Potential Impacts
Many ecosystems within the CBS are strongly affected by the local
activities of often-dense coastal human populations. Activities such as
the overexploitation of fisheries, unsustainable coastal development,
a
nd pollution have resulted in the widespread degradation of CBS
ecosystems (Burke et al., 2002, 2011). These influences have combined
with steadily increasing ocean temperature and acidification to drive
major changes to a range of important ecosystems over the past 50
years. Understanding the interactions between climate change and non-
climate change drivers is a central part of the detection and attribution
process within the CBS.
Overall, the CBS warmed by 0.14°C to 0.80°C from 1950 to 2009 (Table
30-1), although changes within the Gulf of Mexico/Caribbean Sea sub-
region were not significant (p-value > 0.05) over this period. Key sub-
regions within the CBS such as the Coral Triangle and Western Indian
Ocean warmed by 0.79°C and 0.60°C, respectively, from 1950 to 2009
(Table 30-1). Rates of SLR vary from decreasing sea levels (–5 to –10
mm yr
–1
) to low (2 to 3 mm yr
–1
, Caribbean) to very high (10 mm yr
–1
,
Southeast Asia; Figure 30-5) rates of increase. Ocean acidification also
varies from region to region (Figure SM30-2), and is influenced by
oceanographic and coastal processes, which often have a large human
component.
30.5.4.1.1. Bohai/Yellow Sea/East China Sea
The Bohai Sea, Yellow Sea, and the East China Sea (ECS) are shallow
marginal seas along the edge of the northwest Pacific that are strongly
influenced by the Kuroshio Current (Matsuno et al., 2009), the East
Asian Monsoon (EAM), and major rivers such as the Yellow (Huang He)
and Yangtze (Changjiang) Rivers. Upwelling of the Kuroshio sub-surface
waters provides abundant nutrients that support high levels of primary
productivity (Wong et al., 2000, 2001). The ecosystems of the ECS are
heavily affected by human activities (e.g., overfishing and pollution),
which tend to compound the influence and consequences of climate
change.
SST within the ECS has increased rapidly since the early 1980s (high
confidence; Lin et al., 2005; Jung, 2008; Cai et al., 2011; Tian et al.,
2012). The largest increases in SST have occurred in the ECS in winter
(1.96°C, 1955–2005) and in the Yellow Sea in summer (1.10°C, 1971–
2006; Cai et al., 2011). These changes in SST are closely linked to a
weakening of the EAM (e.g., Cai et al., 2006, 2011; Tang et al., 2009) and
increasing warmth of the Kuroshio Current (Qi et al., 2010; Zhang et al.,
2011; Wu et al., 2012). At the same time, dissolved O
2
has decreased (Lin
et al., 2005; Jung, 2008; Qi et al., 2010), with an associated increase in
the extent of the hypoxic areas in coastal areas of the Yellow Sea/ECS
(Jung, 2008; Tang, 2009; Ning et al., 2011).
Primary productivity, biomass yields, and fish capture rates have
experienced large changes within the ECS over the past decades (limited
evidence, medium agreement; low confidence; Tang et al., 2003; Lin et
al., 2005; Tang, 2009). Fluctuations in herring abundance appear to
closely track SST shifts within the Yellow Sea (Tang, 2009). For plankton
and fish species, the proportions of warm-water species relative to
warm-temperate species in the Changjiang River Estuary (extending to
the southern Taiwan Strait) have changed over past decades (Zhang et
al., 2005; Ma et al., 2009; Lin and Yang, 2011). Northward shifts in catch
distribution for some pelagic fish species in Korean waters were driven,
in part, by warming SST (medium confidence; Jung et al., 2014). The
1687
The Ocean Chapter 30
30
f
requency of harmful algal blooms and blooms of the giant jellyfish
Nemopilema nomurai in the offshore area of the ECS have increased
and have been associated with ocean warming and other factors such
as eutrophication (Ye and Huang, 2003; Tang, 2009; Cai and Tan, 2010).
Although attribution of these changes to anthropogenic climate change
is complicated by the increasing influence of non-climate-related human
activities, many of these changes are consistent with those expected as
SST increases.
30.5.4.1.2. South China Sea
The South China Sea (SCS) is surrounded by continental areas and
includes large numbers of islands, and is connected to the Pacific, ECS,
and Sulu Sea by straits such as the Luzon and Taiwan Strait. The region
is greatly influenced by cyclones/typhoons, and by the Pearl, Red, and
Mekong Rivers. The region has a distinct seasonal circulation and is
greatly influenced by the southwest monsoon (in summer), the Kuroshio
Current, and northeast monsoon (in winter). The SCS includes significant
commercial fisheries areas and includes coral reefs, mangroves, and
seagrass beds.
The surface waters of the SCS have been warming steadily from 1945
to 1999 with the annual mean SST in the central SCS increasing by
0.92°C (1950–2006; Cai et al., 2009), a rate similar to that observed for
the entire Indo-Pacific/Southeast Asian CBS from 1950 to 2009 (0.80°C;
Table 30-1). Significant freshening in the SCS intermediate layer since
the 1960s has been observed (Liu et al., 2007). The temperature change
of the upper layers of the SCS has made a significant contribution to
sea level variation, which is heterogeneous in space and time (Li et al.,
2002; Cheng and Qi, 2007; Liu et al., 2007).
Identifying the extent to which climate change is influencing the SCS is
difficult due to confounding non-climate change factors and their
interactions (e.g., local human pollution, over-exploitation together with
“natural climate variability such as EAM, ENSO, and PDO). Changing
sea temperatures have influenced the abundance of phytoplankton,
benthic biomass, cephalopod fisheries, and the size of demersal trawl
catches in the northern SCS observed over the period 1976–2004
(limited evidence, medium agreement; Ning et al., 2009). Coral reefs
and mangroves are degrading rapidly as a result of both climate change
and non-climate change-related factors (very likely; Box CC-CR; Chen
et al., 2009; China-SNAP, 2011; Zhao et al., 2012). Mass coral bleaching
and mortality of coral reefs within the SCS were triggered by elevated
temperatures in 1998 and 2007 (Yu et al., 2006; Li et al., 2011).
Conversely, warming enabled the establishment of a high-latitude, non-
carbonate, coral community in Daya Bay in northern SCS, although this
community has recently degraded as a result of increasing anthropogenic
stresses (Chen et al., 2009; Qiu et al., 2010).
30.5.4.1.3. Southeast Asian Seas
The Southeast Asian Seas (SAS) include an archipelago of diverse islands
that interact with the westward flow of the North Equatorial Current
and the Indonesian Throughflow (Figure 30-1a). A large part of this
region is referred to as the “Coral Triangle” (Veron et al., 2009). The
w
orld’s most biologically diverse marine area, it includes parts of
Malaysia, Indonesia, the Philippines, Timor Leste, the Solomon Islands,
and Papua New Guinea. SST increased significantly from 1985 to 2006
(Peñaflor et al., 2009; McLeod et al., 2010), although with considerable
spatial variation. Trends examined over longer periods (1950–2009)
show significant warming (+0.80°C, p-value 0.05; Table 30-1). The
sea level is rising by up to 10 mm yr
–1
in much of this region (Church et
al., 2004, 2006; Green et al., 2010). Like other tropical areas in the world,
coral reefs within SAS have experienced periods of elevated temperature,
which has driven several mass coral bleaching and mortality events
since the early 1980s (high confidence; Hoegh-Guldberg et al., 2009;
McLeod et al., 2010; Figure 30-10a). The most recent occurred during
warm conditions in 2010 (Krishnan et al., 2011). These changes are the
result of increasing ocean temperatures and are very likely to be a
consequence of anthropogenic climate change (high confidence; Box
CC-CR; WGI AR5 Section 10.4.1). Although calcification rates of some
key organisms (e.g., reef-building corals; Tanzil et al., 2009) have slowed
over the past 2 decades, it is not possible to conclude that the changes
are due to ocean acidification. While a large part of the decline in coral
reefs has been due to increasing local stresses (principally destructive
fishing, declining water quality, and over-exploitation of key reef species),
projected increases in SST represent a major challenge for these valuable
ecosystems (high agreement; Burke et al., 2002; Burke and Maidens,
2004).
30.5.4.1.4. Arabian Sea and Somali Current
The Arabian Sea and Somali Current are relatively productive ocean
areas, being strongly influenced by upwelling and the monsoonal system.
Wind-generated upwelling enhances primary production in the western
Arabian Sea (Prakash and Ramesh, 2007). Several key fisheries within
this region are under escalating pressure from both fishing and climate
change. SST increased by 0.18°C and 0.26°C in the Arabian Sea and
Somali Current, respectively, from 1982 to 2006 (HadSST2; Rayner et
al., 2003; Belkin, 2009), which is consistent with the overall warming
of the Western Indian Ocean portion of the CBS from 1950 to 2009
(0.60°C; Table 30-1). Salinity of surface waters in the Arabian Sea
increased by 0.5 to 1.0% over the past 60 years (Figure 30-6c), due to
increased evaporation from warming seas and contributions from the
outflows of the saline Red Sea and Arabian Gulf. As in other tropical
sub-regions, increasing sea temperatures have increased the frequency
of mass coral bleaching and mortality within this region (Wilkinson and
Hodgson, 1999; Goreau et al., 2000; Wilkinson, 2004).
The aragonite saturation horizon in both the Arabian Sea and Bay of
Bengal is now 100 to 200 m shallower than in preindustrial times as a
result of ocean acidification (medium confidence; Feely et al., 2004).
Shoaling of the aragonite saturation horizon is likely to affect a range
of organisms and processes, such as the depth distribution of pteropods
(zooplankton) in the western Arabian Sea (medium confidence; Hitchcock
et al., 2002; Mohan et al., 2006). More than 50% of the area of OMZs
in the world’s oceans occur in the Arabian Sea and Bay of Bengal and
long-term measurements reveal that O
2
concentrations are declining in
this region (high confidence; Helly and Levin, 2004; Karstensen et al.,
2008; Stramma et al., 2010; Section 30.3.2.3). The information regarding
the consequences of climate change within this region is undeveloped
1688
Chapter 30 The Ocean
30
a
nd suggests that important physical, chemical, and biological responses
to climate change need to be the focus of further investigation.
30.5.4.1.5. East Africa coast and Madagascar
The Western Indian Ocean strongly influences the coastal conditions
associated with Kenya, Mozambique, Tanzania, Madagascar, La Réunion,
Mayotte, and three archipelagos (Comoros, Mauritius, and the Seychelles).
Sea temperatures in the Western Indian Ocean have increased by 0.60°C
over 1950–2009 (high confidence; p-value 0.05; Table 30-1), increasing
the frequency of positive thermal anomalies that have triggered mass
coral bleaching and mortality events across the region over the past
2 decades (high confidence; Baker et al., 2008; Nakamura et al., 2011;
Box CC-HS). Trends in changes in SST and surface salinity vary with
location along the East African coastline, with faster rates at higher
latitudes (Figure 30-2). Periods of heat stress over the past 20 years have
triggered mass coral bleaching and mortality on coral reef ecosystems
within this region (McClanahan et al., 2007, 2009a,b,c; Ateweberhan
and McClanahan, 2010; Ateweberhan et al., 2011). Steadily increasing
sea temperatures have also produced anomalous growth rates in long-
lived corals such as Porites (high confidence; McClanahan et al., 2009b).
Differences in the susceptibility of reef-building corals to stress from
rising sea temperatures has also resulted in changes to the composition
of coral (high confidence; p-value ≤ 0.05; McClanahan et al., 2007) and
benthic fish communities (high confidence; p-value ≤ 0.05; Graham et
al., 2008; Pratchett et al., 2011a). These changes are very likely to alter
species composition and potentially the productivity of coastal fisheries
(robust evidence, high agreement; high confidence; Jury et al., 2010),
although there may be a significant lag between the loss of coral
communities and the subsequent changes in the abundance and
community structure of fish populations (p-value 0.05; Graham et al.,
2007). Some of these potential changes can be adverted or reduced by
interventions such as the establishment of marine protected areas and
changes to fishing management (McClanahan et al., 2008; Cinner et al.,
2009; Jury et al., 2010; MacNeil et al., 2010).
30.5.4.1.6. Gulf of Mexico and Caribbean Sea
The Gulf of Mexico and Caribbean Sea form a semi-contained maritime
province within the Western Atlantic. These areas are dominated by a
range of activities including mineral extraction, fishing, and tourism, which
provide employment and opportunity for almost 75 million people who
live in coastal areas of the USA, Mexico, and a range of other Caribbean
nations (Adams et al., 2004). The Gulf of Mexico and Caribbean Sea
have warmed by 0.31°C and 0.50°C, respectively, from 1982 to 2006
(very likely; Belkin, 2009). Warming trends are not significant from
1950 to 2009 (Table 30-1), which may be partly due to spatial variability
in warming patterns (Section 30.5.3.1). The Caribbean region has
experienced a sustained decrease in aragonite saturation state from
1996 to 2006 (very likely; Gledhill et al., 2008). Sea levels within the
Gulf of Mexico and Caribbean Sea have increased at the rate of 2 to 3
mm yr
–1
from 1950 to 2000 (Church et al., 2004; Zervas, 2009).
Understanding influences of climate change on ocean ecosystems in
this region is complicated by the confounding influence of growing
h
uman populations and activities. The recent expansion of the seasonal
hypoxic zone, and the associated “dead zone, in the Gulf of Mexico
has been attributed to nitrogen inputs driven by land management
(Turner and Rabalais, 1994; Donner et al., 2004) and changes to river
flows, wind patterns, and thermal stratification of Gulf waters (high
confidence; Justić et al., 1996, 2007; Levin et al., 2009; Rabalais et al.,
2009). The increases in coastal pollution and fishing have potentially
interacted with climate change to exacerbate impacts on marine
ecosystems within this region (Sections 5.3.4, 29.3). These changes have
often been abrupt and non-linear (Taylor et al., 2012).
A combination of local and global disturbances has driven a large-scale
loss of reef-building corals across the Caribbean Sea since the late 1970s
(high confidence; Hughes, 1994; Gardner et al., 2003). Record thermal
stress in 2005 triggered the largest mass coral bleaching and mortality
event on record for the region, damaging coral reefs across hundreds of
square kilometers in the eastern Caribbean Sea (high confidence; Donner
et al., 2007; Eakin et al., 2010). Although conditions in 2010 were milder
than in 2005, elevated temperatures still occurred in some parts of the
Caribbean (Smith et al., 2013). Increasing temperatures in the Caribbean
have also been implicated in the spread of marine diseases (Harvell et
al., 1999, 2002, 2004) and some introduced species (likely; Firth et al.,
2011). As in other sub-regions, pelagic fish species are sensitive to
changes in sea temperature and modify their distribution and abundance
accordingly (Muhling et al., 2011). Fish and invertebrate assemblages
in the Gulf of Mexico have shifted deeper in response to SST warming
over 1970s–2011 (medium confidence; Pinsky et al., 2013).
Coral ecosystems in the Caribbean Sea are at risk from ocean acidification
(very likely; Albright et al., 2010; Albright and Langdon, 2011), although
impacts have yet to be observed under field conditions. Ocean acidification
may also be altering patterns of fish recruitment to coral reefs, although
direct evidence for how this has affected Caribbean species is lacking
(low confidence; Dixson et al., 2008, 2010; Munday et al., 2009).
30.5.4.2. Key Risks and Vulnerabilities
Worldwide, 850 million people live within 100 km of tropical coastal
ecosystems such as coral reefs and mangroves deriving multiple benefits
including food, coastal protection, cultural services, and income from
industries such as fishing and tourism (Burke et al., 2011). Marine
ecosystems within the CBS are sensitive to increasing sea temperatures
(Figure 30-10), although detection and attribution are complicated by the
significant influence and interaction with non-climate change stressors
(water quality, over-exploitation of fisheries, coastal degradation; Box
CC-CR). Warming is likely to have changed the primary productivity of
ocean waters, placing valuable ecosystems and fisheries within the ECS
at risk (low to medium confidence). Other risks include the expansion
of hypoxic conditions and associated dead zones in many parts of the
CBS. Given the consequences for coastal ecosystems and fisheries,
these changes are very likely to increase the vulnerability of coastal
communities throughout the CBS.
Sea temperatures are increasing within many parts of CBS ecosystems
(1950–2009; Table 30-1), and will continue to do so over the next few
decades and century. Sea temperatures are projected to change by
1689
The Ocean Chapter 30
30
(b) Mass coral bleaching: Degree Heating Month (DHM) ≥1
(a)
(c) Mass coral mortality: Degree Heating Month (DHM) ≥5
1.0
0.5
0.0
1.0
0.5
0.0
1.0
0.5
0.0
1.0
0.5
0.0
1.0
0.5
0.0
1.0
0.5
0.0
1.0
0.5
0.0
0.0 0.0
1.0
0.5
1.0
1.01.0
1.0 1.0
0.5
1.0
0.5
0.0
1.0
0.5
0.0
1.0
0.5
0.0
1900 2000 2100 1900 2000
2100
Proportion of coral grids with DHM of 1 and above
Proportion of coral grids with DHM of 5 and above
1900 2000 2100 1900 2000 2100
1900 2000 21001900 2000 2100
1900 2000 21001900 2000 2100
1900 2000 21001900 2000 2100
1900 2000 2100
1900 2000 2100
Western
Indian Ocean
Eastern
Indian Ocean
Western
Pacific Ocean
Eastern
Pacific Ocean
Caribbean and
Gulf of Mexico
Coral Triangle and
Southeast Asia
Western
Indian Ocean
Eastern
Indian Ocean
Western
Pacific Ocean
Eastern
Pacific Ocean
Caribbean and
Gulf of Mexico
Coral Triangle and
Southeast Asia
Historic
RCP2.6
RCP4.5
RCP6.0
RCP8.5
Figure 30-10 | Annual maximum proportions of reef pixels with Degree Heating Months (DHM, Donner et al., 2007) for each of the six coral regions (a, Figure 30-4b)—(b) DHM ≥1 (used for projecting the incidence of coral bleaching;
Strong et al., 1997, 2011) and (c) DHM ≥5 (associated with bleaching followed by significant mortality; Eakin et al., 2010)—for the period 1870–2009 using the Hadley Centre Interpolated sea surface temperature 1.1 (HadISST1.1) data
set. The black line on each graph is the maximum annual area value for each decade over the period 1870–2009. This value is continued through 2010–2099 using Coupled Model Intercomparison Project Phase 5 (CMIP5) data and splits
into the four Representative Concentration Pathways (RCP2.6, 4.5, 6.0, and 8.5). DHM were produced for each of the four RCPs using the ensembles of CMIP models. From these global maps of DHM, the annual percentage of grid cells
with DHM ≥1 and DHM ≥5 were calculated for each coral region. These data were then grouped into decades from which the maximum annual proportions were derived. The plotted lines for 2010–2099 are the average of these maximum
proportion values for each RCP. Monthly sea surface temperature anomalies were derived using a 1985–2000 maximum monthly mean climatology derived in the calculations for Figure 30-4. This was done separately for HadISST1.1, the
CMIP5 models, and each of the four RCPs, at each grid cell for every region. DHMs were then derived by adding up the monthly anomalies using a 4-month rolling sum. Figure SM30-3 presents past and future sea temperatures for the six
major coral reef provinces under historic, un-forced, RCP4.5 and RCP8.5 scenarios.
Western Pacific Eastern Pacific
Caribbean and Gulf of Mexico
Western Indian Ocean Eastern Indian Ocean
Coral Triangle
The location of shallow-water coral reef cells
Caribbean and
Gulf of Mexico
Caribbean and
Gulf of Mexico
Coral Triangle and
Southeast Asia
Coral Triangle and
Southeast Asia
Eastern
Indian Ocean
Eastern
Indian Ocean
Eastern
Pacific Ocean
Eastern
Pacific Ocean
Western
Indian Ocean
Western
Indian Ocean
Western
Pacific Ocean
Western
Pacific Ocean
1690
Chapter 30 The Ocean
30
0
.34°C to 0.50°C over the near term (2010–2039) and by 0.23°C to
0.74°C over the long term (2010–2099) under the lowest RCP scenario
(RCP2.6). Under BAU (RCP8.5), CBS sea temperatures are projected to
increase by 0.62°C to 0.85°C over the near term and 2.44°C to 3.32°C
over the long term (Table SM30-4). Given the large-scale impacts (e.g.,
mass coral bleaching and mortality events) that have occurred in response
to much smaller changes in the past over CBS regions (0.14°C to 0.80°C
from 1950–2009; Table 30-1), the projected changes of 2.44°C to
3.32°C over 2010–2099 are very likely to have large-scale and negative
consequences for the structure and function of many CBS ecosystems
(virtually certain), especially given the observed sensitivity of coral reefs
to relatively small increases in temperature over the past 3 decades
(Hoegh-Guldberg, 1999; Eakin et al., 2010; Lough, 2012).
It is very likely that coral-dominated reef ecosystems within the CBS
(and elsewhere) will continue to decline and will consequently provide
significantly less ecosystem goods and services for coastal communities if
sea temperatures increase by more than 1°C above current temperatures
(Box CC-CR; Figure 30-10). Combining the known sensitivity of coral
reefs within the Caribbean and Coral Triangle sub-regions (Strong et al.,
1997, 2011; Hoegh-Guldberg, 1999), with the exposure to higher
temperatures that are projected under medium (RCP4.5) to high (RCP8.5)
scenarios, reveals that both coral reef-rich regions are virtually certain
to experience levels of thermal stress (DHM 1) that cause coral
bleaching every 1 to 2 years by the mid- to late part of this century
(robust evidence, high agreement; very high confidence; Figures 30-4b,c,
30-10, 30-12, SM30-3; van Hooidonk et al., 2013). The frequency of mass
mortality events (DHM ≥ 5; Figure 30-10a,b,c) also increases toward a
situation whereevents that occur every 1 to 2 years by the mid- to late
part of this century under low to high climate change scenarios (robust
evidence, high agreement; very high confidence; Hoegh-Guldberg, 1999;
Donner et al., 2005; Frieler et al., 2012). Mass mortality events that
affect coral reefs will result in changes to community composition in
the near term (2010–2039; Berumen and Pratchett, 2006; Adjeroud et
al., 2009) and a continuing downward trend in coral cover in the longer
term (Gardner et al., 2003; Bruno and Selig, 2007; Baker et al., 2008).
It is virtually certain that composition of coral reef fish populations
(Graham et al., 2007; Pratchett et al., 2008, 2011a,b) will change. The
productivity of many fisheries will decrease (limited evidence, medium
agreement) as waters warm, acidify, and stratify, and as crucial habitat,
such as coral reefs, degrade (low confidence). These changes are very
likely to increase the vulnerability of millions of people who live in
coastal communities and depend directly on fisheries and other goods
and services provided by ecosystems such as coral reefs (Hoegh-
Guldberg et al., 2009; McLeod et al., 2010).
30.5.5. Eastern Boundary Upwelling Ecosystems
The Eastern Boundary Upwelling Ecosystems (EBUE) include the California,
Peru/Humboldt, Canary/northwest Africa, and Benguela Currents. They
are highly productive sub-regions with rates of primary productivity that
may exceed 1000 g C m
–2
yr
–1
. Although these provinces comprise less
than 2% of the Ocean area, they contribute nearly 7% of marine primary
production (Figure 30-1b) and more than 20% of the world’s marine
capture fisheries (Pauly and Christensen, 1995). Catches in the EBUE are
d
ominated by planktivorous sardine, anchovy, and horse/jack mackerel,
and piscivorous benthic fish such as hake. Nutrient input from upwelling
of cooler waters stimulates primary production that is transferred to
mid and upper trophic levels, resulting in substantial fish, seabird, and
marine mammal populations. As a result, the EBUE are considered
“hotspots” of productivity and biodiversity (Block et al., 2011). The high
level of productivity is a result of large-scale atmospheric pressure
gradients and wind systems that advect surface waters offshore, leading
to the upwelling of cold, nutrient-rich waters from depth (Box CC-UP;
Chavez and Messie, 2009; Chavez et al., 2011). Upwelling waters are
typically low in pH and high in CO
2
, and are likely to continue to enhance
changes in pH and CO
2
resulting from rising atmospheric CO
2
(Feely et
al., 2008; Gruber, 2011).
30.5.5.1. Observed Changes and Potential Impacts
There are extensive studies of the coupled climate-ecosystem dynamics
of individual EBUE (e.g., California Current). Decadal variability poses
challenges to the detection and attribution of changes within the EBUE
to anthropogenic climate change, although there are a number of long-
term studies that have been able to provide insight into the patterns
of change and their causes. Like other ocean sub-regions, EBUE are
projected to warm under climate change, with increased stratification
and intensified winds as westerly winds shift poleward (likely). However,
cooling has also been predicted for some EBUE, resulting from the
intensification of wind-driven upwelling (Bakun, 1990). The California
and Canary Currents have warmed by 0.73°C and 0.53°C (very likely;
p-value 0.05, 1950–2009; Table 30-1), respectively, while no significant
trend was detected in the sea surface temperatures of the Benguela
(p-value = 0.44) and Humboldt Currents (p-value = 0.21) from 1950 to
2009 (Table 30-1). These trends match shorter-term trends for various
EBUE using Pathfinder version 5 data (Demarcq, 2009). These differences
are likely to be the result of differences in the influence of long-term
variability and the specific responses of coastal wind systems to warming,
although an analysis of wind data over the same period did not pick
up clear trends (low confidence, with respect to long-term wind trends;
Demarcq, 2009; Barton et al., 2013).
How climate change will influence ocean upwelling is central to resolving
ecosystem and fishery responses within each EBUE. There is considerable
debate, however, as to whether or not climate change will drive an
intensification of upwelling (e.g., Bakun et al., 2010; Narayan et al., 2010;
Barton et al., 2013) in all regions. This debate is outlined in Box CC-UP.
EBUE are also areas of naturally low pH and high CO
2
concentrations
due to upwelling, and consequently may be vulnerable to ocean
acidification and its synergistic impacts (Barton et al., 2012). A full
understanding of the consequences of ocean acidification for marine
organisms and ecosystems is discussed elsewhere (Boxes CC-OA, CC-UP;
Sections 6.2, 6.3.2; Kroeker et al., 2013; WGI AR5 Section 6.4).
30.5.5.1.1. Canary Current
Part of the North Atlantic STG, the Canary Current extends from northern
Morocco southwestward to the North Atlantic Equatorial Current. It is
linked with the Portugal Current (which is sometimes considered part
1691
The Ocean Chapter 30
30
o
f the Canary Current) upstream. The coastal upwelling system,
however, is limited to a narrow belt along the Saharan west coast to
the coast of Guinea, with the most intense upwelling occurring centrally,
along the coasts of Mauritania (15°N to 20°N) and Morocco (21°N to
26°N). Total fish catches, comprising mainly coastal pelagic sardines,
sardinellas, anchovies, and mackerel, have fluctuated around 2 million
tonnes yr
–1
since the 1970s (www.seaaroundus.org/lme/27.aspx).
Contrasting with the other EBUE, fishing productivity is modest, probably
partly due to the legacy of uncontrolled fishing in the 1960s (Arístegui
et al., 2009).
Most observations suggest that the Canary Current has warmed since
the early 1980s (Arístegui et al., 2009; Belkin, 2009; Demarcq, 2009;
Barton et al., 2013), with analysis of HadISST1.1 data from 1950 to 2009
indicating warming of 0.53°C from 1950–2009 (p-value ≤ 0.05; Table
30-1). Gómez-Gesteira et al. (2008) suggest a 20 and 45% decrease in
the strength of upwelling in winter and summer, respectively, from 1967
to 2006, consistent with a decrease in wind strength and direction over
the past 60 years. More recently, Barton et al. (2013) show no clear
increasing or decreasing trend in wind strength over the past 60 years,
and a lack of agreement among wind trends and variability from
different wind products (e.g., Pacific Fisheries Environmental Laboratory
(PFEL), International Comprehensive Ocean-Atmosphere Data Set
(ICOADS), Wave- and Anemometer-based Sea Surface Wind (WASWind)).
Barton et al. (2013) present no evidence for changes in upwelling
intensity, with the exception of upwelling off northwest Spain, where
winds are becoming slightly less favorable. Alteration of wind direction
and strength influences upwelling and hence nutrient concentrations;
however, nutrient levels can also change in response to other variables
such as the supply of iron-laden dust from the Sahara (Alonso-Pérez et
al., 2011). There is medium evidence and medium agreement that
primary production in the Canary Current has decreased over the past
2 decades (Arístegui et al., 2009; Demarcq, 2009), in contrast to the
nearby upwelling region off northwest Spain where no significant trend
was observed (Bode et al., 2011). Satellite chlorophyll records (Sea-
viewing Wide Field-of-view Sensor (SeaWiFS), Moderate Resolution
Imaging Spectrometer (MODIS)) are relatively short, making it difficult
to distinguish the influence of warming oceans from longer term patterns
of variability (Arístegui et al., 2009; Henson et al., 2010). Changing
temperature has resulted in changes to important fisheries species. For
example, Mauritanian waters have become more suitable as feeding
and spawning areas for some fisheries species (e.g., Sardinella aurita)
as temperatures increased (Zeeberg et al., 2008). Clear attribution of
these changes depends on the linkage between the Azores High and
global temperature, and on longer records for both physical and
biological systems, as pointed out for data sets in general (Arístegui et
al., 2009; Henson et al., 2010).
30.5.5.1.2. Benguela Current
The Benguela Current originates from the eastward-flowing, cold South
Atlantic Current, flows northward along the southwest coast of Africa,
and is bounded north and south by the warm-water Angola and Agulhas
Currents, respectively. Upwelling is strongest and most persistent toward
the center of the system in the Lüderitz-Orange River upwelling cell
(Hutchings et al., 2009). Fish catch reached a peak in the late 1970s of
2
.8 million tonnes yr
–1
(
www.seaaroundus.org/lme/29/1.aspx), before
declines in the northern Benguela, due to overfishing and inter-decadal
environmental variability, resulted in a reduced catch of around 1 million
tonnes yr
1
(present) (Cury and Shannon, 2004; Heymans et al., 2004;
Hutchings et al., 2009). Offshore commercial fisheries currently comprise
sardine, anchovy, horse mackerel, and hake, while the inshore artisanal
and recreational fisheries comprise a variety of fish species mostly
caught by hook and line.
Most research on the Benguela Current has focused on fisheries and
oceanography, with little emphasis on climate change. As with the other
EBUE, strong interannual and inter-decadal variability in physical
oceanography make the detection and attribution of biophysical trends
to climate change difficult. Nevertheless, the physical conditions of the
Benguela Current are highly sensitive to climate variability over a range
of scales, especially to atmospheric teleconnections that alter local wind
stress (Hutchings et al., 2009; Leduc et al., 2010; Richter et al., 2010;
Rouault et al., 2010). Consequently, there is medium agreement, despite
limited evidence (Demarcq, 2009), that upwelling intensity and associated
variables (e.g., temperature, nutrient, and O
2
concentrations) from the
Benguela system will change as a result of climate change (Box CC-UP).
The temperature of the surface waters of the Benguela Current did not
increase from 1950 to 2009 (p-value > +0.05; Table 30-1), although
shorter records show an decrease in the south-central Benguela Current
(0.35°C to 0.55°C per decade; Rouault et al., 2010) or an increase for
the whole Benguela region (0.24°C; Belkin, 2009). These differences
between short versus long records indicate the substantial influence of
long-term variability on the Benguela system (Belkin, 2009). Information
on other potential consequences of climate change within the Benguela
system is sparse. SLR is similar to the global mean, although it has not
been measured rigorously within the Benguela (Brundrit, 1995; Veitch,
2007). Although upwelling water in the northern and southern portions
of the Benguela Current exhibits elevated and suppressed partial pressure
of CO
2
, respectively (Santana-Casiano et al., 2009), the consequences
of changing upwelling intensity remain poorly explored with respect to
ocean acidification. Finally, although periodic hypoxic events in the
Benguela system are largely driven by natural advective processes, these
may be exacerbated by future climate change (Monteiro et al., 2008;
Bakun et al., 2010).
Despite its apparent sensitivity to environmental variability, there is
limited evidence of ecological changes in the Benguela Current EBUE
due to climate change (Poloczanska et al., 2013). For example, pelagic
fish (Roy et al., 2007), benthic crustaceans (Cockcroft et al., 2008), and
seabirds (Crawford et al., 2008) have demonstrated general eastward
range shifts around the Cape of Good Hope. Although these may be
associated with increased upwelling along the South African south
coast, specific studies that attribute these changes to anthropogenic
climate change are lacking. Trawl surveys of demersal fish and cephalopod
species showed consistently predictable “hotspots” of species richness
over a 20- to 30-year study period (the earliest surveys since 1984 off
South Africa) that were associated with greater depths and cooler
bottom waters (Kirkman et al., 2013). However, major changes in the
structure and function of the demersal community have been shown in
some parts of the Benguela Current EBUE in response to environmental
change, for example, due predominantly to fishing pressure in the 1960s
1692
Chapter 30 The Ocean
30
a
nd environmental forcing in the early 2000s in the southern Benguela
(Howard et al., 2007); therefore, changes driven by climate change may
eventually affect the persistence of these biodiversity hotspots (Kirkman
et al., 2013).
30.5.5.1.3. California Current
The California Current spans approximately 23° of latitude from central
Baja California, Mexico, to central British Columbia, Canada, linking the
North Pacific Current (West Wind Drift) with the North Equatorial and
Kuroshio Currents to form the North Pacific Gyre. High productivity driven
by advective transport and upwelling (Hickey, 1979; Chelton et al., 1982;
Checkley and Barth, 2009; Auad et al., 2011) supports well-studied
ecosystems and fisheries. Fish catches have been approximately 0.6 million
tonnes yr
–1
since 1950 (www.seaaroundus.org/lme/3.aspx), which makes
it the lowest catch of the four EBUE. The ecosystem supports the foraging
and reproductive activities of 2 to 6 million seabirds from around 100
species (Tyler et al., 1993). Marine mammals are diverse and relatively
abundant, including recovering populations of humpback whales,
among other species (Barlow et al., 2008).
The average temperature of the California Current warmed by 0.73°C
from 1950 to 2009 (p-value 0.05; Table 30-1) and by 0.14°C to 0.80°C
from 1985 to 2007 (Demarcq, 2009). Like other EBUE, the California
Current is characterized by large-scale interannual and inter-decadal
climate-ecosystem variability (McGowan et al., 1998; Hare and Mantua,
2000; Chavez et al., 2003; Checkley and Barth, 2009). During an El Niño,
coastally trapped Kelvin waves from the tropics deepen the thermocline,
thereby severely reducing upwelling and increasing ocean temperatures
from California to Washington (Peterson and Schwing, 2003; King et al.,
2011). Atmospheric teleconnections to the tropical Pacific alter wind
stress and coastal upwelling. Therefore, the ENSO is intimately linked
with Bakun’s (1990) upwelling intensification hypothesis (Box CC-UP).
Inter-decadal variability in the California Current stems from variability
in the Pacific-North America pattern (Overland et al., 2010), which is
influenced by the PDO (Mantua et al., 1997; Peterson and Schwing,
2003) and the NPGO (Di Lorenzo et al., 2008). The major effects of the
PDO and NPGO appear north of 39°N (Di Lorenzo et al., 2008; Menge
et al., 2009).
There is robust evidence and medium agreement that the California
Current has experienced a decrease in the number of upwelling events
(23 to 40%), but an increase in duration of individual events, resulting in
an increase of the overall magnitude of upwelling events from 1967 to
2010 (high confidence; Demarcq, 2009; Iles et al., 2012). This is consistent
with changes expected under climate change yet remains complicated
by the influence of decadal-scale variability (low confidence; Iles et al.,
2012). Oxygen concentrations have also undergone large and consistent
decreases from 1984 to 2006 throughout the California Current, with
the largest relative decreases occurring below the thermocline (21% at
300 m). The hypoxic boundary layer (<60 µmol kg
–1
) has also shoaled
by up to 90 m in some regions (Bograd et al., 2008). These changes are
consistent with the increased input of organic carbon into deeper layers
from enhanced upwelling and productivity, which stimulates microbial
activity and results in the drawdown of O
2
(likely, Bakun et al., 2010; but
see also McClatchie et al., 2010; Koslow et al., 2011; WGI AR5 Section
3
.8.3). These changes are likely to have reduced the available habitat
for key benthic communities as well as fish and other mobile species
(Stramma et al., 2010). Increasing microbial activity will also increase the
partial pressure of CO
2
, decreasing the pH and carbonate concentration
of seawater. Together with the shoaling of the saturation horizon, these
changes have increased the incidence of low O
2
and low pH water
flowing onto the continental shelf (high confidence; 40 to 120 m;
Feely et al., 2008), causing problems for industries such as the shellfish
aquaculture industry (Barton et al., 2012).
30.5.5.1.4. Humboldt Current
The Humboldt Current is the largest of the four EBUE, covering an area
larger than the other three combined. It comprises the eastern edge
of the South Pacific Gyre, linking the northern part of the Antarctic
Circumpolar Current with the Pacific South Equatorial Current. Although
the primary productivity per unit area is modest compared to that of
the other EBUE, the total Humboldt Current system has very high levels
of fish production. Current catches are in line with a long-term average
(since the 1960s) of 8 million tonnes yr
1
(www.seaaroundus.org/lme/
13/1.aspx), although decadal-scale variations range from 2.5 to 13
million tonnes yr
1
. While anchovies currently contribute 80% of the
total catch, they alternate with sardines on a multi-decadal scale, with
their dynamics mediated by the approach and retreat of subtropical
waters to and from the coast (Alheit and Bakun, 2010). This variability
does not appear to be changing due to anthropogenic climate change.
Thus, from the late 1970s to the early 1990s, sardines were more
important (Chavez et al., 2003). The other major commercial fish species
are jack mackerel among the pelagic fish and hake among the demersal
fish.
The Humboldt Current EBUE did not show an overall warming trend in
SST over the last 60 years (p-value > 0.05; Table 30-1), which is consistent
with other data sets (1982–2006, HadISST1.1: Belkin, 2009; 1985–2007,
Pathfinder: Demarcq, 2009). Wind speed has increased in the central
portions of the Humboldt Current, although wind has decreased in its
southern and northern sections (Demarcq, 2009). The lack of a consistent
warming signal may be due to the strong influence of adjacent ENSO
activity exerting opposing drivers on upwelling and which, if they
intensify, would decrease temperatures (limited evidence, medium
agreement). Similar to the Canary Current EBUE, however, there was a
significant increase in the temperatures of the warmest month of the
year over the period 1950–2009 (p-value ≤ 0.05; Table 30-1).
Primary production is suppressed during warm El Niño events and
amplified during cooler La Niña phases, these changes then propagate
through to higher trophic levels (Chavez et al., 2003; Tam et al., 2008;
Taylor et al., 2008). However, in addition to trophic changes, there is
also a direct thermal impact on organisms, which varies depending on
the thermal adaptation window for each species (high confidence). A
37-year zooplankton time series for the coast of Peru showed no
persistent trend in abundance and diversity (Ayón et al., 2004), although
observed shifts coincided with the shifts in the regional SST. As for other
EBUE, there is lack of studies that have rigorously attempted to detect
and attribute changes to anthropogenic climate change, although at
least two studies (Mendelssohn and Schwing, 2002; Gutiérrez et al.,
1693
The Ocean Chapter 30
30
2
011) provide additional evidence that the northern Humboldt Current
has cooled (due to upwelling intensification) since the 1950s, a trend
matched by increasing primary production. This is not entirely consistent
with the lack of significant change over the period 1950–2009 (p-value
> 0.05; Table 30-1). Nevertheless, these relationships are likely to be
complex in their origin, especially in their sensitivity to the long-term
changes associated with ENSO and PDO, and the fact that areas within
the Humboldt Current EBUE may be showing different behaviors.
30.5.5.2. Key Risks and Vulnerabilities
EBUE are vulnerable to changes that influence the intensity of currents,
upwelling, and mixing (and hence changes in SST, wind strength and
direction), as well as O
2
content, carbonate chemistry, nutrient content,
and the supply of organic carbon to deep offshore locations (robust
evidence, high agreement; high confidence). The extent to which any
particular EBUE is vulnerable to these factors depends on location
(Figure 3 from Gruber, 2011) and other factors such as alternative
sources of nutrient input and fishing pressure (Bakun et al., 2010). This
complex interplay between regional and global drivers means that our
understanding of how factors such as upwelling within the EBUE will
respond to further climate change is uncertain (Box CC-UP; Rykaczewski
and Dunne, 2010).
In the GCM ensembles examined (Table SM30-3), modest rates of
warming (0.22°C to 0.93°C) occur within the four EBUEs in the near
term. Over 2010–2099, however, EBUE SSTs warm by 0.07°C to 1.02ºC
under RCP2.6, and 2.52°C to 3.51ºC under RCP8.5 (Table SM30-4).
These high temperatures have the potential to increase stratification of
the water column and substantially reduce overall mixing in some areas.
In contrast, the potential strengthening of coastal wind systems would
intensify upwelling and stimulate primary productivity through the
increased injection of nutrients into the photic zone of the EBUE (Box
CC-UP). Garreaud and Falvey (2009) explored how wind stress along
t
he South American coast would change by 2100 under SRES B2 and
A2 scenarios. Using an ensemble of 15 GCMs, southerly wind systems
upwelling increased along the subtropical coast of South America,
extending and strengthening conditions for upwelling.
Changes in the intensity of upwelling within the EBUE will drive
fundamental changes to the abundance, distribution, and viability of
resident organisms, although an understanding of their nature and
direction is limited. In some cases, large-scale decreases in primary
productivity and dependent fisheries are projected to occur for EBUE
ecosystems (Blanchard et al., 2012), while other projections question
the strong connection between primary productivity and fisheries
production (Arístegui et al., 2009). Increased upwelling intensity also
has potential disadvantages. Elevated primary productivity may lead
to decreasing trophic transfer efficiency, thus increasing the amount of
organic carbon exported to the seabed, where it is virtually certain to
increase microbial respiration and hence increase low O
2
stress (Weeks
et al., 2002; Bakun et al., 2010). Increased wind stress may also increase
turbulence, breaking up food concentrations (affecting trophic transfer),
or causing excessive offshore advection, which could remove plankton
from shelf habitats. The central issue for the EBUE is therefore whether
or not upwelling will intensify and, if so, whether the negative
consequences (e.g., reduced O
2
and elevated CO
2
) associated with
upwelling intensification will outweigh potential benefits from increased
primary production and fisheries catch.
30.5.6. Subtropical Gyres
Subtropical gyres (STG) dominate the Pacific, Atlantic, and Indian
Oceans (Figure 30-1a), and consist of large stable water masses that
circulate clockwise (Northern Hemisphere) and anticlockwise (Southern
Hemisphere) due to the Coriolis Effect. The oligotrophic areas at the
core of the STG represent one of the largest habitats on Earth,
contributing 21.2% of ocean primary productivity and 8.3% of the
Frequently Asked Questions
FAQ 30.4 | Will climate change increase the number of “dead zones” in the oceans?
Dissolved oxygen is a major determinant of the distribution and abundance of marine organisms. Dead zones are
persistent hypoxic conditions where the water doesn’t have enough dissolved oxygen to support oxygen-dependent
marine species. These areas exist all over the world and are expanding, with impacts on coastal ecosystems and
fisheries (high confidence). Dead zones are caused by several factors, particularly eutrophication where too many
nutrients run off coastal cities and agricultural areas into rivers that carry these materials out to sea. This stimulates
primary production, leading to a greater supply of organic carbon, which can sink into the deeper layers of the
ocean. As microbial activity is stimulated, there is a sharp reduction in dissolved oxygen levels and an increased risk
of dead zones (high confidence). Climate change can influence the distribution of dead zones by increasing water
temperature and hence microbial activity, as well as reducing mixing (i.e., increasing layering or stratification) of
the Ocean, thereby reducing mixing of oxygen-rich surface layers into the deeper parts of the Ocean. In other
areas, increased upwelling can lead to stimulated productivity, which can also lead to more organic carbon entering
the deep ocean, where it is consumed, decreasing oxygen levels (medium confidence). Managing local factors such
as the input of nutrients into coastal regions can play an important role in reducing the rate at which dead zones
are spreading across the world’s oceans (high agreement).
1694
Chapter 30 The Ocean
30
g
lobal fish catch (Figure 30-1b; Table SM30-1). A number of small island
nations are found within this region. While many of the observed
changes within these nations have been described in previous chapters
(e.g., Sections 5.3-4, 29.3-5), region-wide issues and consequences are
discussed here due to the strong linkages between ocean and coastal
issues.
30.5.6.1. Observed Changes and Potential Impacts
The central portions of the STG are oligotrophic (Figure SM30-1).
Temperatures within the STG of the North Pacific (NPAC), South Pacific
(SPAC), Indian Ocean (IOCE), North Atlantic (NATL), and South Atlantic
(SATL) have increased at rates of 0.020°C, 0.024°C, 0.032°C, 0.025°C,
and 0.027°C yr
–1
from 1998 to 2010, respectively (Signorini and McClain,
2012). This is consistent with increases observed from 1950 to 2009
(0.25°C to 0.67ºC; Table 30-1). However, differences among studies
done over differing time periods emphasize the importance of long-
term patterns of variability. Salinity has decreased across the North and
South Pacific STG (Figure 30-6c; WGI AR5 Section 3.3.3.1), consistent
with warmer sea temperatures and an intensification of the hydrological
cycle (Boyer, 2005).
The North and South Pacific STG have expanded since 1993 (high
confidence), with these changes likely being the consequence of a
combination of wind forcing and long-term variability (Parrish et al.,
2000; WGI AR5 Section 3.6.3). Chlorophyll levels, as determined by
remote-sensing of ocean color (Box CC-UP), have decreased in the
NPAC, IOCE, and NATL by 9, 12, and 11%, respectively (p-value ≤ 0.5;
Signorini and McClain, 2012) over and above the inherent seasonal and
interannual variability from 1998 to 2010 (Vantrepotte and Mélin, 2011).
Chlorophyll levels did not change in the remaining two gyres (SPAC and
SATL, and confirmed for SPAC by Lee and McPhaden (2010) and Lee et
al. (2010)). Furthermore, over the period 1998–2007, median cell diameter
of key phytoplankton species exhibited statistically significant linear
declines of about 2% in the North and South Pacific, and 4% in the
North Atlantic Ocean (Polovina and Woodworth, 2012). Changes in
chlorophyll and primary productivity in these sub-regions have been noted
before (McClain et al., 2004; Gregg et al., 2005; Polovina et al., 2008)
and are influenced by seasonal and longer-term sources of variability
(e.g., ENSO, PDO; Section 6.3.4; Figure 6-9). These changes represent a
significant expansion of the world’s most unproductive waters, although
caution must be exercised given the limitations of satellite detection
methods (Box CC-PP) and the shortness of records relative to longer-
term patterns of climate variability. There is high confidence that
changes that reduce the vertical transport of nutrients into the euphotic
zone (e.g., decreased wind speed, increasing surface temperatures, and
stratification) will reduce the rate of primary productivity and hence
fisheries.
30.5.6.1.1. Pacific Ocean Subtropical Gyres
Pacific climate is heavily influenced by the position of the Intertropical
Convergence Zone (ITCZ) and the South Pacific Convergence Zone
(SPCZ), which are part of the ascending branch of the Hadley circulation
(WGI AR5 Section 14.3.1). These features are also strongly influenced
b
y interannual to inter-decadal climate patterns of variability including
ENSO and PDO. The current understanding of how ENSO and PDO will
change as average global temperatures increase is not clear (low
confidence; Collins et al., 2010; WGI AR5 Section 12.4.4.2). The position
of both the ITCZ and SPCZ vary seasonally and with ENSO (Lough et
al., 2011), with a northward migration during the Northern Hemisphere
summer and a southward migration during the Southern Hemisphere
summer. These changes, along with the West Pacific Monsoon, determine
the timing and extent of the wet and dry seasons in SPAC and NPAC
sub-regions (Ganachaud et al., 2011). Tropical cyclones are prominent
in the Pacific (particularly the western Pacific), and CBS sub-regions
between 1and 30° north and south of the equator, although the
associated storm systems may occasionally reach higher latitudes.
Spatial patterns of cyclones vary with ENSO, spreading out from the
Coral Sea to the Marquesas Islands during El Niño and contracting back
to the Coral Sea, New Caledonia, and Vanuatu during La Niña (Lough
et al., 2011). Historically, there have been almost twice as many land-
falling tropical cyclones in La Niña as opposed to El Niño years off the
east coast of Australia, with a declining trend in the number of severe
tropical cyclones from 0.45 per year in the early 1870s to 0.17 per year
in recent times (Callaghan and Power, 2011).
The Pacific Ocean underwent an abrupt shift to warmer sea temperatures
in the mid-1970s as a result of both natural (e.g., IPO) and climate forcing
(high confidence; Meehl et al., 2009). This change coincided with changes
to total rainfall, rain days, and dry spells across the Pacific, with the
direction of change depending on the location relative to the SPCZ.
Countries such as the Cook Islands, Tonga, Samoa and American Samoa,
and Fiji tend to experience drought conditions as the SPCZ (with cooler
sea temperatures) moves toward the northeast during El Niño (high
confidence). The opposite is true during La Niña conditions. The
consequences of changing rainfall on the countries of the Pacific STG are
discussed in greater detail elsewhere (Sections 5.4, 29.3; Table 29-1).
Although these changes are due to different phases of long-term
variability in the Pacific, they illustrate the ramifications and sensitivity
of the Pacific to changes in climate change.
Elevated sea temperatures within the Pacific Ocean have increased the
frequency of widespread mass coral bleaching and mortality since the
early 1980s (very high confidence; Hoegh-Guldberg and Salvat, 1995;
Hoegh-Guldberg, 1999; Mumby et al., 2001; Baker et al., 2008; Donner et
al., 2010). There are few, if any, scientific records of mass coral bleaching
and mortality prior to this period (high confidence; Hoegh-Guldberg,
1999). Rates of decline in coral cover on coastal coral reef ecosystems
range between 0.5 and 2.0% per year depending on the location within
the Indo-Pacific region (high confidence; Bruno and Selig, 2007; Hughes
et al., 2011; Sweatman et al., 2011; De’ath et al., 2012). The reasons for
this decline are complex and involve non-climate change-related factors
(e.g., coastal pollution and overfishing) as well as global warming and
possibly acidification. A recent comprehensive analysis of the ecological
consequences of coral bleaching and mortality concluded that “bleaching
episodes have resulted in catastrophic loss of coral reefs in some locations,
and have changed coral community structure in many others, with a
potentially critical influence on the maintenance of biodiversity in the
marine tropics” (high confidence; Baker et al., 2008, p. 435). Increasing
sea levels have also caused changes in seagrass and mangrove systems.
Gilman et al. (2007) found a reduction in mangrove area with SLR, with
1695
The Ocean Chapter 30
30
t
he observed mean landward recession of three mangrove areas over
4 decades being 25, 64, and 72 mm yr
1
, 12 to 37 times faster than the
observed rate of SLR. Significant interactions exist between climate
change and coastal development, where migration shoreward depends
on the extent to which coastlines have been modified or barriers to
successful migration have been established.
Changes in sea temperature also lead to changes in the distribution
of key pelagic fisheries such as skipjack tuna (Katsuwonus pelamis),
yellowfin tuna (Thunnus albacares), big-eye tuna (T. obesus), and South
Pacific albacore tuna (T. alalunga), which make up the majority of key
fisheries in the Pacific Ocean. Changes in distribution and recruitment in
response to changes in sea temperature as result of ENSO demonstrate
the close association of pelagic fish stocks and water temperature. The
shift in habitat for top predators in the northeast Pacific was examined
by Hazen et al. (2012), who used tracking data from 23 marine species
and associated environmental variables to predict changes of up to 35%
in core habitat for these species within the North Pacific. Potential habitats
are predicted to contract for the blue whale, salmon shark, loggerhead
turtle, and blue and mako sharks, while potential habitats for the sooty
shearwater; black -footed albatross; leatherback turtle; white shark;
elephant seal; and albacore, bluefin and yellowfin tuna are predicted to
expand (Hazen et al., 2012). However, expansion of OMZs in the Pacific
STG is predicted to compress habitat (depth) for hypoxia-intolerant
species such as tuna (Stramma et al., 2010, 2012).
Reduction of ocean productivity of the STG (Sarmiento et al., 2004;
Signorini and McClain, 2012) reduces the flow of energy to higher
trophic levels such as those of pelagic fish (Le Borgne et al., 2011). The
distribution and abundance of fisheries stocks such as tuna are also
sensitive to changes in sea temperature, and hence long-term variability
such as ENSO and PDO. The redistribution of tuna in the western central
equatorial region has been related to the position of the oceanic
convergence zones, where the warm pool meets the cooler tongue of
the Pacific. These changes have been reliably reproduced by population
models that use temperature as a driver of the distribution and abundance
of tuna (Lehodey et al., 1997, 2006). Projections of big-eye tuna (T. obesus)
distributions under SRES A2 show an improvement in spawning and
feeding habitats by 2100 in the eastern tropical Pacific and declines in
the western tropical Pacific, leading to an eastern displacement of tuna
stocks (Lehodey et al., 2008, 2010b).
30.5.6.1.2. Indian Ocean Subtropical Gyre
Like the Pacific Ocean, the Indian Ocean plays a crucial role in global
weather patterns, with teleconnections throughout Africa, Australasia,
Asia, and the Americas (e.g., Clark et al., 2000; Manhique et al., 2011;
Meehl and Arblaster, 2011; Nakamura et al., 2011). Increasing sea level,
temperature, storm distribution and intensity, and changing seawater
chemistry all influence the broad range of physical, chemical, and
biological aspects of the Indian Ocean. Coral reef ecosystems in the
Indian Ocean gyre system were heavily affected by record positive sea
temperature anomalies seen in the Southern Hemisphere between
February to April 1998 (robust evidence, high agreement; high confidence;
Ateweberhan et al., 2011). Coral cover across the Western Indian Ocean
declined by an average of 37.7% after the 1998 heat stress event
(
Ateweberhan et al., 2011). Responses to the anomalously warm
conditions in 1998 varied between sub-regions, with the central Indian
Ocean islands (Maldives, Seychelles, Chagos, and Lakshadweep)
experiencing major decreases in coral cover directly after the 1998 event
(from 40 to 53% coral cover in 1977–1997 to 7% in 1999–2000; high
confidence; Ateweberhan et al., 2011). Coral reefs lining the islands of
southern India and Sri Lanka experienced similar decreases in coral cover
(45%, 1977–1997 to 12%, 1999–2000). Corals in the southwestern
Indian Ocean (Comoros, Madagascar, Mauritius, Mayotte, Réunion, and
Rodrigues) showed less impact (44%, 1977–1997 to 40%, 1999–2000).
Recovery from these increases in mortality has been variable, with sites
such as those around the central Indian Ocean islands exhibiting fairly
slow recovery (13% by 2001–2005) while those around southern India
and Sri Lanka are showing much higher rates (achieving a mean coral
cover of 37% by 2001–2005; Ateweberhan et al., 2011). These changes
to the population size of key reef-building species will drive major
changes in the abundance and composition of fish populations in
coastal areas, and affect other ecosystem services that are important
for underpinning tourism and coastal protection (medium confidence;
Box CC-CR).
Fisheries that exploit tuna and other large pelagic species are very
valuable to many small island states within the Indian Ocean. As with
Pacific fisheries, the distribution and abundance of large pelagic fish
in the Indian Ocean is greatly influenced by sea temperature. The
anomalously high sea temperatures of 1997–1998 (leading to a
deepening of the mixed layer in the west and a shoaling in the east)
coincided with anomalously low primary production in the Western
Indian Ocean and a major shift in tuna stocks (high confidence; Menard
et al., 2007; Robinson et al., 2010). Fishing grounds in the Western Indian
Ocean were deserted and fishing fleets underwent a massive shift toward
the eastern basin, which was unprecedented for the tuna fishery (high
confidence). As a result of these changes, many countries throughout
the Indian Ocean lost significant tuna-related revenue (Robinson et al.,
2010). In 2007, tuna fishing revenue was again reduced by strong
surface warming and deepening of the mixed layer, and associated with
a modest reduction in primary productivity in the west. These trends
highlight the overall vulnerability of tuna fishing countries in the Indian
Ocean to climate variability, a situation similar to that in the other major
oceans of the world.
30.5.6.1.3. Atlantic Ocean Subtropical Gyres
SST has increased within the two STG of the Atlantic Ocean over the
last 2 decades (Belkin, 2009; Signorini and McClain, 2012). Over longer
periods of time (1950–2009), trends in average temperature are not
significant for the North Atlantic STG (p-value > 0.05) while they remain
so for the South Atlantic STG (very likely; 0.08°C per decade, p-value ≤
0.05; Table 30-1). In both cases, however, temperatures in the coolest
and warmest months increased significantly (Table 30-1). The difference
between these studies (i.e., over 10 to 30 years vs. 60 years) emphasizes
the importance of long-term patterns of variability in the North Atlantic
region. Variability in SST at a period of about 60 to 80 years is associated
with the Atlantic Multi-decadal Oscillation (AMO; Trenberth and Shea,
2006). Sea surface temperatures influence hurricane activity (very likely)
with recent record SST associated with record hurricane activity in 2005
1696
Chapter 30 The Ocean
30
i
n the Atlantic (Trenberth and Shea, 2006) and mass coral bleaching
and mortality in the eastern Caribbean (high confidence; Eakin et al.,
2010). In the former case, analysis concluded that 0.1°C of the SST
anomaly was attributable to the state of the AMO while 0.45°C was due
to ocean warming as a result of anthropogenic influences (Trenberth and
Shea, 2006).
These changes have influenced the distribution of key fishery species as
well the ecology of coral reefs in Bermuda (Wilkinson and Hodgson,
1999; Baker et al., 2008) and in the eastern Caribbean (Eakin et al., 2010).
Small island nations such as Bermuda depend on coral reefs for fisheries
and tourism and are vulnerable to further increases in sea temperature
that cause mass coral bleaching and mortality (high confidence; Box
CC-CR; Figure 30-10). As with the other STG, phytoplankton communities
and pelagic fish stocks are sensitive to temperature changes that have
occurred over the past several decades. Observation of these changes
has enabled development of models that have a high degree of accuracy
in projecting the distribution and abundance of these elements within
the Atlantic region in general (Cheung et al., 2011).
30.5.6.2. Key Risks and Vulnerabilities
SSTs of the vast STGs of the Atlantic, Pacific, and Indian Oceans are
increasing, which is very likely to increase stratification of the water
column. In turn, this is likely to reduce surface concentrations of nutrients
and, consequently, primary productivity (medium confidence; Box CC-PP).
Warming is projected to continue (Table SM30-4), with substantial
increases in the vulnerability and risk associated with systems that have
been observed to change so far (high confidence; Figure 30-12). Under
RCP2.6, the temperatures of the STG are projected to increase by 0.17°C
to 0.56°C in the near term (over 2010–2039) and between –0.03°C to
0.90°C in the long term (over 2010–2099) (Table SM30-4). Under
RCP8.5, however, surface temperatures of the world’s STG are projected
to be 0.45°C to 0.91°C warmer in the near term and 1.90°C to 3.44°C
warmer in the long term (Table SM30-4). These changes in temperature
are very likely to increase water column stability, reduce the depth of the
mixed layer, and influence key parameters such as nutrient availability
and O
2
concentrations. It is not clear as to how longer-term sources of
variability such as ENSO and PDO will change (WGI AR5 Sections 14.4,
14.7.6) and ultimately influence these trends.
The world’s most oligotrophic ocean sub-regions are likely to continue to
expand over coming decades, with consequences for ecosystem services
such as gas exchange, fisheries, and carbon sequestration. Polovina et
al. (2011) explored this question for the North Pacific using a climate
model that included a coupled ocean biogeochemical component to
investigate potential changes under an SRES A2 scenario (~RCP6.0 to
RCP8.5; see also Figure 1.5 from Rogelj et al., 2012). Model projections
indicated the STG expanding by approximately 30% by 2100, driven
by the northward drift of the mid-latitude westerlies and enhanced
stratification of the water column. The expansion of the STG occurred
at the expense of the equatorial upwelling and other regions within the
North Pacific. In the North Pacific STG, the total primary production is
projected to decrease by 10 to 20% and large fish catch by 19 to 29%
by 2100 under SRES A2 (Howell et al., 2013; Woodworth-Jefcoats et al.,
2013). However, our understanding of how large-scale eddy systems
w
ill change in a warming world is incomplete, as are the implications
for primary productivity of these large and important systems (Boxes
CC-PP, CC-UP).
Understanding how storm frequency and intensity will change represents
a key question for many countries and territories within the various STG.
Projections of increasing sea temperature are likely to change the
behavior of tropical cyclones. At the same time, the maximum wind
speed and rainfall associated with cyclones is likely to increase, although
future trends in cyclones and severe storms are very likely to vary from
region to region (WGI AR5 Section 14.6). Patterns such as “temporal
clustering” can have a strong influence on the impact of tropical cyclones
on ecosystems such as coral reefs (Mumby et al., 2011), although how
these patterns will change within all STG is uncertain at this point.
However, an intensifying hydrological cycle is expected to increase
precipitation in many areas (high confidence; WGI AR5 Sections 2.5,
14.2), although longer droughts are also expected in other STG (medium
confidence). Changes in the hydrological cycle impact coastal ecosystems,
increasing damage through coastal flooding and physical damage from
storm waves (Mumby et al., 2011). Improving our understanding of how
weather systems associated with features such as the SPCZ (WGI AR5
Section 14.3.1) will vary is critical to climate change adaptation of a large
number of nations associated with the STG. Developing an understanding
of how ocean temperature, climate systems such as the SPCZ and ITCZ,
and climate change and variability (e.g., ENSO, PDO) interact will be
essential in this regard. For example, variability in the latitude of the
SPCZ is projected to increase, possibly leading to more extreme events
in Pacific Island countries (Cai et al., 2012).
The consequences of projected sea temperatures on the frequency of
coral bleaching and mortality within key sub-regions of the STG are
outlined in Box CC-CR and Figures 30-10 and SM30-3. As with other
sub-regions (particularly CBS, STG, and SES) dominated by coral reefs,
mass coral bleaching and mortality becomes an annual risk under all
scenarios, with mass mortality events beginning to occur every 1 to 2
years by 2100 (virtually certain; Box CC-CR; Figures 30-10, SM30-3).
Coral-dominated reef ecosystems (areas with more than 30% coral
cover) are very likely to disappear under these circumstances by the mid
part of this century (van Hooidonk et al., 2013). The loss of substantial
coral communities has implications for the three-dimensional structure
of coral reefs (Box CC-CR) and the role of the latter as habitat for
organisms such as fish (Hoegh-Guldberg, 2011; Hoegh-Guldberg et al.,
2011a; Pratchett et al., 2011a; Bell et al., 2013b).
The consequences of increasing sea temperature can be exacerbated
by increasing ocean acidification, with potential implications for reef
calcification (medium confidence; Kleypas et al., 1999; Hoegh-Guldberg
et al., 2007; Doney et al., 2009), reef metabolism and community
calcification (Dove et al., 2013), and other key ecological processes
(Pörtner et al., 2001, 2007; Munday et al., 2009). Ocean pH within
the STG will continue to decrease as atmospheric CO
2
increases,
bringing pH within the STG to 7.9 and 7.7 at atmospheric concentrations
of 450 ppm and 800 ppm, respectively (Figure SM30-2a; Box CC-OA).
Aragonite saturation states will decrease to around 1.6 (800 ppm) and
3.3 (450 ppm; Figure SM30-2b). Decreasing carbonate ion concentrations
and saturation states pose serious risks to other marine calcifiers
such as encrusting coralline algae, coccolithophores (phytoplankton),
1697
The Ocean Chapter 30
30
a
nd a range of benthic invertebrates (Doney et al., 2009; Feely et al.,
2009).
Increasing sea temperatures and sea level are also likely to influence
other coastal ecosystems (e.g., mangroves, seagrass meadows) in the
Pacific, although significant gaps and uncertainties exist (Section 29.3.1.2;
Waycott et al., 2007, 2011). Many of the negative consequences for
coral reefs, mangroves, and seagrass meadows are likely to have
negative consequences for dependent coastal fisheries (through habitat
destruction) and tourism industries (medium confidence; Bell et al.,
2011a, 2013a; Pratchett et al., 2011a,b).
Populations of key large pelagic fish are projected to move many
hundreds of kilometers east of where they are today in the Pacific
STG (high confidence; Lehodey et al., 2008, 2010a, 2011, 2013), with
implications for income, industry, and food security across multiple
Pacific Island nations (high confidence; Cheung et al., 2010; McIlgorm
et al., 2010; Bell et al., 2011b, 2013a; Section 7.4.2; Tables 29-2, 29-4).
These predictions of species range displacements, contractions, and
expansions in response to anticipated changes in the Ocean (Box CC-MB)
present both a challenge and an opportunity for the development of
large-scale management strategies to preserve these valuable species.
Our understanding of the consequences of reduced O
2
for pelagic fish
populations is not clear, although there is high agreement on the potential
physiological outcomes (Section 6.3.3). Those species that are intolerant
to hypoxia, such as skipjack and yellowfin tuna (Lehodey et al., 2011),
will have their depth range compressed in the Pacific STG, which will
increase their vulnerability to fisheries and reduce overall fisheries
habitat and productivity (medium confidence; Stramma et al., 2010, 2011).
Despite the importance of these potential changes, our understanding
of the full range of consequences is limited at this point.
30.5.7. Deep Sea (>1000 m)
Assessments of the influence of climate change on the Deep Sea (DS)
are challenging because of difficulty of access and scarcity of long-term,
comprehensive observations (Smith, Jr. et al., 2009). The size of this habitat
is also vast, covering well over 54% of the Earths surface and stretching
from the top of the mid-oceanic ridges to the bottom of deep ocean
trenches (Smith, Jr. et al., 2009). The fossil record in marine sediments
reveals that the DS has undergone large changes in response to climate
change in the past (Knoll and Fischer, 2011). The paleo-skeletal record
shows that it is the rate, not just the magnitude, of climate change
(temperature, O
2
, and CO
2
) that is critical to marine life in DS. The current
rate of change in key parameters very likely exceeds that of other major
events in Earth history. Two primary time scales are of interest. The first
is the slow rate (century-scale) of ocean circulation and mixing, and
consequently the slow rate at which DS ecosystems experience physical
climate change. The second is the rapid rate at which organic matter
enters the deep ocean from primary productivity generated at the surface
of the Ocean, which represents a critical food supply to DS animals
(Smith, Jr. and Kaufmann, 1999; Smith, Jr. et al., 2009). It can also represent
a potential risk in some circumstances where the flux of organic carbon
into the deep ocean, coupled with increased sea temperatures, can lead
to anoxic areas (dead zones) as metabolism is increased and O
2
decreased
(Chan et al., 2008; Stramma et al., 2010).
30.5.7.1. Observed Changes and Potential Impacts
The greatest rate of change of temperature is occurring in the upper 700
m of the Ocean (very high confidence; WGI AR5 Section 3.2), although
smaller yet significant changes are occurring at depth. The DS environment
is typically cold (~–0.5°C to 3°C; Smith et al., 2008), although abyssal
temperatures in the SES can be higher (e.g., Mediterranean DS ~12°C;
Danovaro et al., 2010). In the latter case, DS organisms can thrive in these
environments as well, illustrating the variety of temperature conditions
that differing species of abyssal life have adapted to. Individual species,
however, are typically constrained within a narrow thermal and O
2
-
demand window of tolerance (Pörtner, 2010) and therefore it is likely
that shifts in the distribution of DS species and regional extinctions will
occur. Warming over multiple decades has been observed below 700 m
(Levitus et al., 2005, 2009), with warming being minimal at mid-range
depths (2000 to 3000 m), and increasing toward the sea floor in some
sub-regions (e.g., Southern Ocean; WGI AR5 Chapter 3). For the deep
Atlantic Ocean, the mean age of deep waters (mean time since last
exposure to the atmosphere) is approximately 250 years; the oldest deep
waters of the Pacific Ocean are >1000 years old. The patterns of ocean
circulation are clearly revealed by the penetration of tracers and the
signal of CO
2
released from burning fossil fuel penetrating into the abyss
(Sabine et al., 2004). It will take many centuries for full equilibration of
deep ocean waters and their ecosystems with recent planetary warming
and CO
2
levels (Wunsch and Heimbach, 2008).
Temperature accounts for approximately 86% of the variance in the
export of organic matter to the DS (medium confidence; Laws et al.,
2000). Consequently, upper ocean warming will reduce the export of
organic matter to the DS (medium confidence), potentially changing the
distribution and abundance of DS organisms and associated food webs,
and ecosystem processes (Smith, Jr. and Kaufmann, 1999). Most organic
matter entering the DS is recycled by microbial systems at relatively
shallow depths (Buesseler et al., 2007), and at rates that are temperature
dependent. Upper ocean warming will increase the rate of sub-surface
decomposition of organic matter (high confidence), thus intensifying
the intermediate depth OMZs (Stramma et al., 2008, 2010) and reducing
food supply to the abyssal ocean.
Particulate organic carbon is exported from the surface to deeper layers
of the Ocean (>500 m) with an efficiency of between 20 and 50%
(Buesseler et al., 2007), much of it being recycled by microbes before it
reaches 1000 m (Smith, Jr. et al., 2009). The export of organic carbon is
dependent on surface net primary productivity, which is likely to vary
(Box CC-PP), influencing the supply of food to DS (Laws et al., 2000;
Smith et al., 2008). Warming of intermediate waters will also increase
respiration at mid-water depths, reducing the flux of organic carbon. Our
understanding of other components of DS ecosystems is also relatively
poor. For example, there is limited evidence and limited agreement as
to how ocean warming and acidification are likely to affect ecosystems
such as those associated with hydrothermal vents (Van Dover, 2012).
Oxygen concentrations are decreasing in the DS (Stramma et al., 2008;
Helm et al., 2011a). Although the largest signals occur at intermediate
water depths < 1000 m (Nakanowatari et al., 2007; Whitney et al.,
2007; Falkowski et al., 2011), some waters >1000 m depth are also
experiencing a decline (Jenkins, 2008). The quantity of dissolved O
2
1698
Chapter 30 The Ocean
30
t
hroughout the Ocean will be reduced with warming due to direct
effects on solubility (high confidence), with these effects being widely
distributed (Shaffer et al., 2009). It is also virtually certain that metabolic
rates of all animals and microbial respiration rates will increase with
temperature (Brown et al., 2004). Thus, increased microbial activity
and reduced O
2
solubility at higher temperatures will have additive
consequences for the decline of O
2
(high confidence) even in the
DS. The DS waters are relatively well oxygenated owing to the higher
solubility of O
2
in colder waters and the low supply rate of organic
matter to great depths. The availability of oxygen to marine animals is
governed by a combination of concentration, temperature, pressure,
and related properties such as diffusivity. Analysis by Hofmann et al.
(2013) reveals that the supply potential of oxygen to marine animals
in cold deep waters is similar to that at much shallower depths (very
high confidence).
Anthropogenic CO
2
has penetrated to at least 1000 m in all three
ocean basins (particularly the Atlantic; Doney et al., 2009). Further
declines of calcite and aragonite in already under-saturated DS water
will presumably decrease biological carbonate structure formation and
increase dissolution, as has happened many times in Earth’s past (high
confidence; Zeebe and Ridgwell, 2011). Some cold-water corals (reported
down to 3500 m) already exist in waters under-saturated with respect
to aragonite (Lundsten et al., 2009). Although initial investigations
suggested that ocean acidification (reduced by 0.15 and 0.30 pH units)
would result in a reduction in the calcification rate of deep water corals
(30 and 56%, respectively), accumulating evidence shows that ocean
acidification may have far less impact than previously anticipated on
the calcification of some deep water corals (limited evidence, medium
agreement; low confidence) although it may reduce important habitats
given that dead unprotected coral mounds are likely to dissolve in
under-saturated waters (Thresher et al., 2011; Form and Riebesell, 2012;
Maier et al., 2013).
30.5.7.2. Key Risks and Vulnerabilities
Rising atmospheric CO
2
poses a risk to DS communities through increasing
temperature, decreasing O
2
and pH, and changing carbonate chemistry
(high confidence; Keeling et al., 2010). Risks associated with the DS have
implications for the Ocean and planet given the high degree of inherent
dependency and connectivity. The resulting changes to the flow of organic
carbon to some parts of the DS (e.g., STG) are very likely to affect DS
ecosystems (medium confidence; Smith et al., 2008). As with the Ocean
generally, there is a need to fill in the substantial gaps that exist in our
knowledge and understanding of the world’s largest habitat and its
responses to rapid anthropogenic climate change.
30.5.8. Detection and Attribution of Climate Change
Impacts with Confidence Levels
The analysis in this chapter and elsewhere in AR5 has identified a wide
range of physical, chemical, and ecological components that have
changed over the last century (Box CC-MB). Figure 30-11 summarizes
a number of examples from the Ocean as a region together with the
degree of confidence in both the detection and attribution steps. For
o
cean warming and acidification, confidence is very high that changes
are being detected and that they are due to changes to the atmospheric
GHG content. There is considerable confidence in both the detection
(very high confidence) and attribution (high confidence) of mass coral
bleaching and mortality, given the well-developed understanding of
environmental processes and physiological responses driving these events
(Box CC-CR; Section 6.3.1). For other changes, confidence is lower, either
because detection of changes has been difficult, or monitoring programs
are not long established (e.g., field evidence of declining calcification),
or because detection has been possible but models are in conflict (e.g.,
wind-driven upwelling). The detection and attribution of recent changes
is discussed in further detail in Sections 18.3.3-4.
30.6. Sectoral Impacts, Adaptation,
and Mitigation Responses
Human welfare is highly dependent on ecosystem services provided by
the Ocean. Many of these services are provided by coastal and shelf
areas, and are consequently addressed in other chapters (e.g., Sections
5.4.3, 7.3.2.4, 22.3.2.3). Oceans contribute provisioning (e.g., food, raw
materials; see Section 30.6.2.1), regulating (e.g., gas exchange, nutrient
recycling, carbon storage, climate regulation, water flux), supporting
(e.g., habitat, genetic diversity), and cultural (e.g., recreational, religious)
services (MEA, 2005; Tallis et al., 2013). The accumulating evidence
indicating that fundamental ecosystem services within the Ocean are
shifting rapidly should be of major concern, especially with respect to
the ability of regulating and supporting ecosystem services to underpin
current and future human population demands (Rockström et al., 2009;
Ruckelshaus et al., 2013). Discussion here is restricted to environmental,
economic, and social sectors that have direct relevance to the Ocean—
namely natural ecosystems, fisheries and aquaculture, tourism, shipping,
oil and gas, human health, maritime security, and renewable energy.
The influences of climate change on Ocean sectors will be mediated
through simultaneous changes in multiple environmental and ecological
variables (see Figure 30-12), and the extent to which changes can be
adapted to and/or risks mitigated (Table 30-3). Both short- and longer-
term adaptation is necessary to address impacts arising from warming,
even under the lowest stabilization scenarios assessed.
Sectoral approaches dominate resource use and management in the
Ocean (e.g., shipping tends to be treated in isolation from fishing within
an area), yet cumulative and interactive effects of individual stressors
are known to be ubiquitous and substantial (Crain et al., 2008). Climate
change consistently emerges as a dominant stressor in regional- to
global-scale assessments, although land-based pollution, commercial
fishing, invasive species, coastal habitat modification, and commercial
activities such as shipping all rank high in many places around the
world (e.g., Sections 5.3.4, 30.5.3-4; Halpern et al., 2009, 2010). Such
cumulative effects pose challenges to managing for the full suite of
stressors to marine systems, but also present opportunities where
mitigating a few key stressors can potentially improve overall ecosystem
condition (e.g., Halpern et al., 2010; Kelly et al., 2011). The latter has often
been seen as a potential strategy for reducing negative consequences
of climate impacts on marine ecosystems by boosting ecosystem
resilience, thus buying time while the core issue of reducing GHG
emissions is tackled (West et al., 2009).
1699
The Ocean Chapter 30
30
30.6.1. Natural Ecosystems
Adaptation in natural ecosystems may occur autonomously, such as
tracking shifts in species’ composition and distributions (Poloczanska
et al., 2013), or engineered by human intervention, such as assisted
dispersal (Section 4.4.2.4; Hoegh-Guldberg et al., 2008). Currently,
adaptation strategies for marine ecosystems include reducing additional
stressors (e.g., maintaining water quality, adapting fisheries management)
and maintaining resilience ecosystems (e.g., Marine Protected Areas),
and are moving toward whole-of-ecosystem management approaches.
Coral reefs, for example, will recover faster from mass coral bleaching
and mortality if healthy populations of herbivorous fish are maintained
(medium confidence; Hughes et al., 2003), indicating that reducing
overfishing will help maintain coral-dominated reef systems while the
international community reduces the emissions of GHGs to stabilize
global temperature and ocean chemistry.
Approaches such as providing a formal valuation of ecological services
from the Ocean have potential to facilitate adaptation by underpinning
more effective governance, regulation, and ocean policy while at the
same time potentially improving management of these often vulnerable
services through the development of market mechanisms and incentives
(Beaudoin and Pendleton, 2012). Supporting, regulating, and cultural
ecosystem services tend to transcend the immediate demands placed
on provisioning services and are difficult to value in formal economic
terms owing to their complexity, problems such as double counting,
and the value of non-market goods and services arising from marine
ecosystems generally (Fu et al., 2011; Beaudoin and Pendleton, 2012).
“Blue Carbon” is defined as the organic carbon sequestered by marine
ecosystems such as phytoplankton, mangrove, seagrass, and salt marsh
ecosystems (Laffoley and Grimsditch, 2009; Nellemann et al., 2009). In
this respect, Blue Carbon will provide opportunities for both adaptation
to, and mitigation of, climate change if key uncertainties in inventories,
methodologies, and policies for measuring, valuing, and implementing
Blue Carbon strategies are resolved (McLeod et al., 2011). Sediment
surface levels in vegetated coastal habitats can rise several meters over
thousands of years, building carbon-rich deposits (Brevik and Homburg,
4
86 7
5
9
1
2 3
I J
A
B C
E F
HG
D
Very low
Very low
Low Medium
Degree of confidence in detection
High
Very high
Low Medium
Degree of confidence in attribution
High Very high
1
. Increased ocean temperature and decreased pH (WGI Sections 3.2,
3
.8, 10.4; WGII Sections 30.3.1–2, 30.5)
2. Decrease in dissolved oxygen (WGI Sections 3.8, 10.4, 30.3.2)
3
. Increase in temperature extremes (WGI Sections 30.3.2, WGII
S
ections 30.3.1, 30.5.3–4)
4. Increased water column stratification and reduced ocean mixing
and ventilation (WGI Section 3.8; WGII Section 30.3.1)
5. Decreased Deep Sea dissolved oxygen and pH (Sections 30.3.2,
30.5.7)
6
. Increased upwelling (Sections 30.5.5, 30.5.2)
7
. Changes in wind and wave stress (Sections 30.3.1, 30.5.1, 30.5.6)
8. Increase in hypoxic areas (Sections 30.5.3–5)
9. Expansion of low-productivity waters (WGI Section 3.8.4; WGII
Section 30.5.6)
Physical and chemical systems
Biological systems
E
quatorial Upwelling Systems (EUS)
S
ubtropical Gyres (STG)
A
Coral bleaching and mortality (Sections 30.5.3–4, 30.5.6, Box
C
C-CR)
B. Changes in distribution of tuna stocks (Section 30.5.6)
C. Poleward expansion of distributions of plankton communities
(Section 30.5.1)
D. Redistribution of marine plants and animals to higher latitudes
(Sections 30.4, 30.5.1–6, Box CC-MB)
E. Increased fisheries catch potential (Section 30.5.1, Box CC-MB)
F. Decreased fisheries catch potential (Sections 30.5.3–4, 30.5.6, Box
CC-MB)
G. Changes to net primary productivity (Box CC-PP)
H. Changes to Deep Sea ecosystems (Section 30. 5.7)
I. Changes to reproduction and migration (Sections 30.4, 30.5.1–6)
J. Decreased calcification (Sections 30.5.3–4, 30.5.6, Box CC-OA)
H
igh-Latitude Spring Bloom Systems (HLSBS)
E
astern Boundary Upwelling Systems (EBUE)
Semi-Enclosed Seas (SES)
Deep Sea (DS)
Coastal Boundary Systems (CBS)
Figure 30-11 | Expert assessment of degree of confidence in detection and attribution of physical and chemical changes (white circles) and ecological changes (dark gray
squares) across sub-regions, as designated in Figure 30-1a, and processes in the Ocean (based on evidence explored throughout Chapter 30 and elsewhere in AR5). Further
explanation of this figure is given in Sections 18.3.3-4 and 18.6.
<1000 mDepth >1000 m
1700
Chapter 30 The Ocean
30
High-Latitude Spring Bloom Systems
Semi-Enclosed Seas
Coastal Boundary Systems Equatorial Upwelling Systems
Subtropical GyresEastern Boundary Upwelling Systems
(1) Expansion of low-
productivity areas as a
consequence of thermal
stratification and changes in
wind stress. (Low)
(2) Northward expansion of
plankton, invertebrate, and
fish communities with
w
arming. Increase in fish
b
iomass at high latitude
f
ringes. (High)
(3) Upwelling, hence productivity, changes as
a result of climate variability, particularly wind
stress variability. (Low)
(
5) Declining dissolved
o
xygen concentrations and
expansion of Oxygen
Minimum Zones (OMZ).
(High)
(6) Decline in dissolved oxygen
through changes in oxygen
solubility (temperature), and
ocean ventilation and circulation.
(Medium)
(7) Expansion of seasonally
hypoxic waters due to thermal
stratification and
e
utrophication. Mass coral
b
leaching events leading to
d
egradation of coral reefs and
loss of associated biodiversity.
(High)
(
8) Shoaling of aragonite
s
aturation horizon reduces
biological calcification and alters
plankton communities. (High)
(9) Mass coral bleaching and
mortality in response to warming
as well as temperature-driven
decline in growth rates of some
corals leading to degradation of
coral reefs and loss of associated
biodiversity. (Very high)
Examples of projected impacts and vulnerabilities associated with climate change in Ocean regions
(4) Spread of tropical species
originating from Indian and Atlantic
Oceans. Increased frequency of mass
m
ortality events of benthic plants
a
nd animals linked to extreme
t
emperature events. (High)
(A) Acidification of seasonally
upwelling waters impacts
shellfish aquaculture.
(Medium)
(B) Increased fish catches at
high-latitude fringes with
economic disruptions and
jurisdictional tensions as
some fish stocks shift
distributions. (Medium)
(C) Thermal stratification
and eutrophication
reduces dissolved
oxygen with impacts on
fish stocks. (Medium)
(D) Sea level rise
modifies coastlines
and increases flooding,
challenging
aquaculture such as
shrimp farms.
(Medium)
(E) Warming leads to
decline in primary
production and
reductions in fish catch.
(Low)
(F) Increase in variability of
upwelling in some EBUEs adds
uncertainty to fisheries
management. (Medium)
(G) Degradation of coral reefs and
associated fish stocks as the extent
and intensity of mass coral
bleaching and mortality increases,
increasing risks to regional food
security. (High)
(H) Temperature-driven shifts
in stocks of large pelagic fish
create winners and losers
among country and island
economies. (High)
Examples of risks to fisheries from observed and projected impacts
A
E
B
B
C
C
D
F
H
H
G
2
2
3
4
5
8
9
6
7
7
1
Figure 30-12 | Top: Examples of projected impacts and vulnerabilities associated with climate change in Ocean sub-regions. Bottom: Examples of risks to fisheries from
observed and projected impacts across Ocean sub-regions. Words in parentheses indicate level of confidence. Details of sub-regions are given in Table 30-1a and Section 30.1.1.
1701
The Ocean Chapter 30
30
2
004; Lo Iacono et al., 2008). The degradation of coastal habitats not
only liberates much of the carbon associated with vegetation loss, but
can also release and oxidize buried organic carbon through erosion of
cleared coastlines (high confidence; Duarte et al., 2005). Combining
data on global area, land use conversion rates, and near-surface carbon
stocks for marshes, mangroves, and seagrass meadows, Pendleton et
al. (2012) revealed that the CO
2
emissions arising from destruction of
these three ecosystems was equivalent to 3 to 19% of the emissions
generated by deforestation globally, with economic damages estimated
to be US$6 to US$42 billion annually. Similarly, Luisetti et al. (2013)
estimate the carbon stock of seagrass and salt marshes in Europe,
representing less than 4% of global carbon stocks in coastal vegetation,
was valued at US$180 million, at EU Allowance price of €8/tCO
2
in June
2012. A reversal of EU Environmental Protection Directives could result
in economic losses of US$1 billion by 2060. Blue Carbon strategies can
also be justified in light of the numerous ecosystem services these
ecosystems provide, such as protection against coastal erosion and
storm damage, and provision of habitats for fisheries species (Section
5.5.7).
30.6.2. Economic Sectors
30.6.2.1. Fisheries and Aquaculture
The Ocean provided 64% of the production supplied by world fisheries
(capture and aquaculture) in 2010, amounting to 148.5 million tonnes
of fish and shellfish (FAO, 2012). This production, valued at US$217.5
billion, supplied, on average, 18.6 kg of protein-rich food per person
to an estimated population of 6.9 billion (FAO, 2012). Marine capture
fisheries supplied 77.4 million tonnes with highest production from the
northwest Pacific (27%), west-central Pacific (15%), northeast Atlantic
(11%), and southeast Pacific (10%) (FAO, 2012). World aquaculture
production (59.9 million tonnes in 2010) is dominated by freshwater
fishes; nevertheless, marine aquaculture supplied 18.1 million tonnes
(30%) (FAO, 2012).
Marine capture fisheries production increased from 16.8 million tonnes
in 1950 to a peak of 86.4 million tonnes in 1996, then declined before
stabilizing around 80 million tonnes (FAO, 2012). The stagnation of marine
capture fisheries production is attributed to full exploitation of around
60% of the worlds marine fisheries and overexploitation of 30%
(estimates for 2009) (FAO, 2012). Major issues for industrial fisheries
include illegal, unreported, and unregulated fishing; ineffective
implementation of monitoring, control, and surveillance; and overcapacity
in fishing fleets (World Bank and FAO, 2008; FAO, 2012). Such problems
are being progressively addressed in several developed and developing
countries (Hilborn, 2007; Pitcher et al., 2009; Worm et al., 2009), where
investments have been made in stock assessment, strong management,
and application of the FAO Code of Conduct for Responsible Fisheries
and the FAO Ecosystem Approach to Fisheries Management.
The significance of marine capture fisheries is illustrated powerfully by
the number of people engaged in marine small-scale fisheries (SSF) in
developing countries. SSF account for around half of the fish harvested
from the Ocean, and provide jobs for more than 47 million people—
about 12.5 million fishers and another 34.5 million people engaged in
p
ost-harvest activities (Mills et al., 2011). SSF are often characterized by
large numbers of politically weak fishers operating from decentralized
localities, with poor governance and insufficient data to monitor catches
effectively (Kurien and Willmann, 2009; Cochrane et al., 2011; Pomeroy
and Andrew, 2011). For these SSF, management that aims to avoid
further depletion of overfished stocks may be more appropriate in
the short-term than management aimed at maximizing sustainable
production. These aims are achieved through adaptive management by
(1) introduction of harvest controls (e.g., size limits, closed seasons and
areas, gear restrictions, and protection of spawning aggregations) to
avoid irreversible damage to stocks in the face of uncertainty (Cochrane
et al., 2011); (2) flexible modification of these controls through
monitoring (Plagányi et al., 2013); and (3) investing in the social capital
and institutions needed for communities and governments to manage
SSF (Makino et al., 2009; Pomeroy and Andrew, 2011).
Changes to ocean temperature, chemistry, and other factors are generating
new challenges for fisheries resulting in loss of coastal and oceanic
habitat (Hazen et al., 2012; Stramma et al., 2012), the movement of
species (Cheung et al., 2011), the spread and increase of disease and
invading species (Ling, 2008; Raitsos et al., 2010; Chan et al., 2011),
and changes in primary production (Chassot et al., 2010). There is
medium evidence and medium agreement that these changes will
change both the nature of fisheries and their ability to provide food and
protein for hundreds of millions of people (Section 7.2.1.2). The risks to
ecosystems and fisheries vary from region to region (Section 7.3.2.4).
Dynamic bioclimatic envelope models under SRES A1B project potential
increases in fisheries production at high latitudes, and potential
decreases at lower latitudes by the mid-21st century (Cheung et al.,
2010; Section 6.5). Overall, warming temperatures are projected to shift
optimal environments for individual species polewards and redistribute
production; however, changes will be region specific (Cheung et al.,
2010; Merino et al., 2012).
Fisheries, in particular shellfish, are also vulnerable to declining pH
and carbonate ion concentrations. As a result, the global production of
shellfish fisheries is likely to decrease (Cooley and Doney, 2009; Pickering
et al., 2011) with further ocean acidification (medium confidence;
Sections 6.3.2, 6.3.5, 6.4.1.1; Box CC-OA). Impacts may be first observed
in EBUE where upwelled water is already relatively low in O
2
and under-
saturated with aragonite (Section 30.5.5). Seasonal upwelling of
acidified waters onto the continental shelf in the California Current
region has recently affected oyster hatcheries along the coast of
Washington and Oregon (Barton et al., 2012; Section 30.5.5.1.1). Whether
declining pH and aragonite saturation due to climate change played a
role is unclear; however, future declines will increase the risk of such
events occurring.
Most marine aquaculture species are sensitive to changing ocean
temperature (Section 6.3.1.4; exposed through pens, cages, and racks
placed directly in the sea, utilization of seawater in land-based tanks,
collection of wild spat) and, for molluscs particularly, changes in carbonate
chemistry (Turley and Boot, 2011; Barton et al., 2012; Section 6.3.2.4).
Environmental changes can therefore impact farm profitability, depending
on target species and farm location. For example, a 1°C rise in SST is
projected to shift production of Norwegian salmonids further north but
may increase production overall (Hermansen and Heen, 2012). Industries
1702
Chapter 30 The Ocean
30
f
or non-food products, which can be important for regional livelihoods
such as Black Pearl in Polynesia, are also affected by rising SST. Higher
temperatures are known to affect the quality of pearl nacre, and can
increase levels of disease in adult oysters (Pickering et al., 2011; Bell et
al., 2013b). Aquaculture production is also vulnerable to extreme events
such as storms and floods (e.g., Chang et al., 2013). Flooding and
inundation by seawater may be a problem to shore facilities on low-
lying coasts. For example, shrimp farming operations in the tropics will
be challenged by rising sea levels, which will be exacerbated by
mangrove encroachment and a reduced ability for thorough-drying of
ponds between crops (Della Patrona et al., 2011).
The impacts of climate change on marine fish stocks are expected to
affect the economics of fisheries and livelihoods in fishing nations
through changes in the price and value of catches, fishing costs, income
to fishers and fishing companies, national labor markets, and industry
re-organization (Sumaila et al., 2011; Section 6.4.1). A study of the
potential vulnerabilities of national economies to the effects of climate
change on fisheries, in terms of exposure to warming, relative importance
of fisheries to national economies and diets, and limited societal capacity
to adapt, concluded that a number countries including Malawi, Guinea,
Senegal, Uganda, Sierra Leone, Mozambique, Tanzania, Peru, Colombia,
Venezuela, Mauritania, Morocco, Bangladesh, Cambodia, Pakistan,
Yemen, and Ukraine are most vulnerable (Allison et al., 2009).
Aquaculture production is expanding rapidly (Bostock et al., 2010) and
will play an important role in food production and livelihoods as the
human demand for protein grows. This may also add pressure on
capture fisheries (FAO, 2012; Merino et al., 2012). Two-thirds of farmed
food fish production (marine and freshwater) is achieved with the use
of feed derived from wild-harvested, small, pelagic fish and shellfish.
Fluctuations in the availability and price of fishmeal and fish oil for
feeds, as well as their availability, pose challenges for the growth of
sustainable aquaculture production, particularly given uncertainties in
changes in EBUE upwelling dynamics to climate change (Section
30.5.5). Technological advances and changes in management such as
increasing feed efficiencies, using alternatives to fishmeal and fish oil,
and farming of herbivorous finfish, coupled with economic and regulatory
incentives, will reduce the vulnerability of aquaculture to the impacts
of climate change on small, pelagic fish abundance (Naylor et al., 2009;
Merino et al., 2010; FAO, 2012).
The challenges of optimizing the economic and social benefits of both
industrial fisheries and SSF and aquaculture operations, which often
already include strategies to adapt to climatic variability (Salinger et
al., 2013), are now made more complex by climate change (Cochrane
et al., 2009; Brander, 2010, 2013). Nevertheless, adaptation options
include establishment of early warning systems to aid decision
making, diversification of enterprises, and development of adaptable
management systems (Chang et al., 2013). Vulnerability assessments
that link oceanographic, biological, and socioeconomic systems can
be applied to identify practical adaptations to assist enterprises,
communities, and households to reduce the risks from climate change
and capitalize on the opportunities (Pecl et al., 2009; Bell et al., 2013b;
Norman-López et al., 2013). The diversity of these adaptation options,
and the policies needed to support them, are illustrated by the examples
in the following subsections.
30.6.2.1.1. Tropical fisheries based on large pelagic fish
Fisheries for skipjack, yellowfin, big-eye, and albacore tuna provide
substantial economic and social benefits to the people of Small Island
Developing States (SIDS). For example, tuna fishing license fees
contribute substantially (up to 40%) to the government revenue of
several Pacific Island nations (Gillett, 2009; Bell et al., 2013b). Tuna
fishing and processing operations also contribute up to 25% of gross
domestic product in some of these nations and employ more than
12,000 people (Gillett, 2009; Bell et al., 2013b). Considerable economic
benefits are also derived from fisheries for top pelagic predators in the
Indian and Atlantic Oceans (FAO, 2012; Bell et al., 2013a). Increasing
sea temperatures and changing patterns of upwelling are projected to
cause shifts in the distribution and abundance of pelagic top predator
fish stocks (Sections 30.5.2, 30.5.5-6), with potential to create “winners
and “losers” among island economies as catches of the transboundary
tuna stocks change among and within their exclusive economic zones
(EEZs; Bell et al., 2013a,b).
A number of practical adaptation options and supporting policies have
been identified to minimize the risks and maximize the opportunities
associated with the projected changes in distribution of the abundant
skipjack tuna in the tropical Pacific (Bell et al., 2011b, 2013a; Lehodey
et al., 2011; Table 30-2). These adaptation and policy options include
(1) full implementation of the regional “vessel day scheme, designed to
distribute the economic benefits from the resource in the face of climatic
variability, and other schemes to control fishing effort in subtropical
areas; (2) strategies for diversifying the supply of fish for canneries in
the west of the region as tuna move progressively east; (3) continued
effective fisheries management of all tuna species; (4) energy efficiency
programs to assist domestic fleets to cope with increasing fuel costs
and the possible need to fish further from port; and (5) the eventual
restructuring of regional fisheries management organizations to help
coordinate management measures across the entire tropical Pacific.
Efforts to ensure provision of operational-level catch and effort data
from all industrial fishing operations will improve models for projecting
redistribution of tuna stocks and quotas under climate change (Nicol
et al., 2013; Salinger et al., 2013). Similar adaptation options and policy
responses are expected to be relevant to the challenges faced by tuna
fisheries in the tropical and subtropical Indian and Atlantic Oceans.
30.6.2.1.2. Small-scale fisheries
Small-scale fisheries (SSF) account for 56% of catch and 91% of people
working in fisheries in developing countries (Mills et al., 2011). SSF are
fisheries that tend to operate at family or community level, have low
levels of capitalization, and make an important contribution to food
security and livelihoods. They are often dependent on coastal ecosystems,
such as coral reefs, that provide habitats for a wide range of harvested
fish and invertebrate species. Despite their importance to many developing
countries, such ecosystems are under serious pressure from human
activities including deteriorating coastal water quality, sedimentation,
ocean warming, overfishing, and acidification (Sections 7.2.1.2, 30.3,
30.5; Box CC-CR). These pressures are translating into a steady decline
in live coral cover, which is very likely to continue over the coming
decades, even where integrated coastal zone management is in place
1703
The Ocean Chapter 30
30
(Sections 30.5.4, 30.5.6). For example, coral losses around Pacific Islands
are projected to be as high as 75% by 2050 (Hoegh-Guldberg et al.,
2011a). Even under the most optimistic projections (a 50% loss of
coral by 2050), changes to state of coral reefs (Box CC-CR; Figures 30-10,
30-12) are very likely to reduce the availability of associated fish and
invertebrates that support many of the SSF in the tropics (high confidence).
In the Pacific, the productivity of SSF on coral reefs has been projected to
decrease by at least 20% by 2050 (Pratchett et al., 2011b), which is also
likely to occur in other coral reef areas globally given the similar and
growing stresses in these other regions (Table SM30-1; Section 30.5.4).
Adaptation options and policies for building the resilience of coral reef
fisheries to climate change suggested for the tropical Pacific include (1)
strengthening the management of catchment vegetation to improve
water quality along coastlines; (2) reducing direct damage to coral reefs;
(3) maintaining connectivity of coral reefs with mangrove and seagrass
habitats; (4) sustaining and diversifying the catch of coral reef fish to
maintain their replenishment potential; and (5) transferring fishing
effort from coral reefs to skipjack and yellowfin tuna resources by
installing anchored fish-aggregating devices (FAD) close to shore (Bell
et al., 2011b; 2013a,b; Table 30-2). These adaptation options and
policies represent a “no regrets” strategy in that they provide benefits
for coral reef fisheries and fishers irrespective of climate change and
ocean acidification.
30.6.2.1.3. Northern Hemisphere HLSBS fisheries
The high-latitude fisheries in the Northern Hemisphere span from around
30/35°N to 60°N in the North Pacific and 80°N in the North Atlantic,
covering a wide range of thermal habitats supporting subtropical/
temperate species to boreal/arctic species. The characteristics of these
Adaptation options Supporting policies
Economic
development
F
ull implementation of the vessel day scheme to control fi shing effort by the
Parties to the Nauru Agreement
a
(W–W)
D
iversifying sources of fi sh for canneries in the region and maintaining trade
agreements, e.g., an economic partnership agreement with the European Union
(
W–W)
Continued conservation and management measures for all species of tuna to
m
aintain stocks at healthy levels and make these valuable species more resilient to
climate change (W–W)
E
nergy effi ciency programs to assist fl eets to cope with oil price rises and minimize
CO
2
emissions and reduce costs of fi shing further afi eld as tuna distributions shift
e
ast (W–W)
Pan-Pacifi c tuna management through merger of the WCPFC and Inter-American
T
ropical Tuna Commission to coordinate management measures across the tropical
Pacifi c (L–W)
S
trengthen national capacity to administer the vessel day scheme.
Adjust national tuna management plans and marketing strategies to provide
exible arrangements to buy and sell tuna.
I
nclude implications of climate change in management objectives of the WCPFC.
Apply national management measures to address climate change effects for
s
ubregional concentrations of tuna in archipelagic waters beyond the mandate of
WCPFC.
R
equire all industrial tuna vessels to provide operational-level catch and effort
data to improve the models for redistribution of tuna stocks during climate change.
Food security
M
anage catchment vegetation to reduce transfer of sediments and nutrients
to coasts to reduce damage to adjacent coastal coral reefs, mangroves, and
s
eagrasses that support coastal fi sheries (W–W).
Foster the care of coral reefs, mangroves, and seagrasses by preventing pollution,
m
anaging waste, and eliminating direct damage to these coastal fi sh habitats
(W–W).
P
rovide for migration of fi sh habitats by prohibiting construction adjacent to
mangroves and seagrasses and installing culverts beneath roads to help the plants
c
olonize landward areas as sea level rises (L–W).
Sustain and diversify catches of demersal coastal fi sh to maintain the
replenishment potential of all stocks (L–W).
Increase access to tuna caught by industrial fl eets through storing and selling tuna
and by-catch landed at major ports to provide inexpensive fi sh for rapidly growing
urban populations (W–W).
Install fi sh aggregating devices close to the coast to improve access to fi sh for rural
communities as human populations increase and demersal fi sh decline (W–W).
Develop coastal fi sheries for small pelagic fi sh species, e.g., mackerel, anchovies,
pilchards, sardines, and scads (W–W?).
Promote simple post-harvest methods, such as traditional smoking, salting, and
drying, to extend the shelf life of fi sh when abundant catches are landed (W–W).
S
trengthen governance for sustainable use of coastal fi sh habitats by (1) building
national capacity to understand the threats of climate change; (2) empowering
c
ommunities to manage fi sh habitats; and (3) changing agriculture, forestry, and
mining practices to prevent sedimentation and pollution.
M
inimize barriers to landward migration of coastal habitats during development of
strategies to assist other sectors to respond to climate change.
A
pply “primary sheries management” to stocks of coastal fi sh and shellfi sh to
maintain their potential for replenishment.
Allocate the necessary quantities of tuna from total national catches to food
security to increase access to fi sh for both urban and coastal populations.
Dedicate a proportion of the revenue from fi shing licences to improve access to
tuna for food security.
Include anchored inshore fi sh aggregating devices as part of national infrastructure
for food security.
Livelihoods
Relocate pearl farming operations to deeper water and to sites closer to coral
reefs and seagrass /algal areas where water temperatures and aragonite saturation
levels are likely to be more suitable for good growth and survival of pearl oysters
and formation of high-quality pearls (L–W).
Raise the walls and fl oor of shrimp ponds so that they drain adequately as sea
level rises (L–W).
Identify which shrimp ponds may need to be rededicated to producing other
commodities (L–W).
Provide incentives for aquaculture enterprises to assess risks to infrastructure so
that farming operations and facilities can be “climate-proofed” and relocated if
necessary.
Strengthen environmental impact assessments for coastal aquaculture activities to
include the additional risks posed by climate change.
Develop partnerships with regional technical agencies to provide support for
development of sustainable aquaculture.
Table 30-2 | Examples of priority adaptation options and supporting policies to assist Pacifi c Island countries and territories to minimize the threats of climate change to
the socioeconomic benefi ts derived from pelagic and coastal fi sheries and aquaculture, and to maximize the opportunities. These measures are classifi ed as win–win (W–W)
adaptations, which address other drivers of the sector in the short term and climate change in the long term, or lose–win (L–W) adaptations, where benefi ts do not exceed costs
in the short term but accrue under longer term climate change (modifi ed from Bell et al., 2013b). WCPFC = Western and Central Pacifi c Fisheries Commission.
a
The Parties to the Nauru Agreement are Federated States of Micronesia, Kiribati, Marshall Islands, Nauru, Palau, Papua New Guinea, Solomon Islands, and Tuvalu.
1704
Chapter 30 The Ocean
30
H
LSBS environments, as well as warming trends, are outlined in Section
30.5.1 and Table 30-1. In part, as a result of 30 years of increase in
temperature (Belkin, 2009; Sherman et al., 2009), there has been an
increase in the size of fish stocks associated with high-latitude fisheries
in the Northern Hemisphere. This is particularly the case for the Norwegian
spring-spawning herring, which has recovered from near-extinction as a
result of overfishing and a cooler climate during the 1960s (Toresen and
Østvedt, 2000). The major components of both pelagic and demersal
high-latitude fish stocks are boreal species located north of 50°N.
Climate change is projected to increase high-latitude plankton production
and displace zooplankton and fish species poleward. As a combined
result of these future changes, the abundance of fish (particularly boreal
species) may increase in the northernmost part of the high-latitude
region (Cheung et al., 2011), although increases will only be moderate
in some areas.
The changes in distribution and migration of pelagic fish shows
considerable spatial and temporal variability, which can increase tensions
among fishing nations. In this regard, tension over the Atlantic mackerel
fisheries has led to what many consider the first climate change-related
conflict between fishing nations (Cheung et al., 2012; Section 30.6.5),
and which has emphasized the importance of developing international
collaboration and frameworks for decision making (Miller et al., 2013;
Sections 15.4.3.3, 30.6.7). The Atlantic mackerel has over the recent
decades been a shared stock between the EU and Norway. However,
the recent advancement of the Atlantic mackerel into the Icelandic EEZ
during summer has resulted in Icelandic fishers operating outside the
agreement between the EU and Norway. Earlier records of mackerel
from the first half of the 20th and second half of the 19th century show,
however, that mackerel was present in Icelandic waters during the
earlier warm periods (Astthorsson et al., 2012). In the Barents Sea, the
northeast Arctic cod, Gadus morhua, reached record-high abundance
in 2012 and also reached its northernmost-recorded distribution (82°N)
(ICES, 2012). A further northward migration is impossible as this would
be into the Deep Sea Polar Basin, beyond the habitat of shelf species.
A further advancement eastwards to the Siberian shelf is, however,
possible. The northeast Arctic cod stock is shared exclusively by Norway
and Russia, and to date there has been a good agreement between
those two nations on the management of the stock. These examples
highlight the importance of international agreements and cooperation
(Table 30-4).
The HLSBS fisheries constitute a large-scale high-tech industry, with
large investments in highly mobile fishing vessels, equipment, and land-
based industries with capacity for adapting fisheries management
and industries for climate change (Frontiers Economics, Ltd., 2013).
Knowledge of how climate fluctuations and change affect the growth,
recruitment, and distribution of fish stocks is presently not incorporated
into fisheries management strategies (Perry et al., 2010). These strategies
are vital for fisheries that hope to cope with the challenges of a changing
ocean environment, and are centrally important to any attempt to
develop ecosystem-based management and sustainable fisheries under
climate change. The large pelagic stocks, with their climate-dependent
migration pattern, are shared among several nations. Developing
equitable sharing of fish quotas through international treaties (Table
30-4) is a necessary adaptation for a sustainable fishery. Factors presently
taken into account in determining the shares of quotas are the historical
f
ishery, bilateral exchanges of quotas for various species, and the time
that stocks are in the various EEZs.
30.6.2.2. Tourism
Tourism recreation represents one of the world’s largest industries,
accounting for 9% (>US$6 trillion) of global GDP and employing more
than 255 million people. It is expected to grow by an average of 4%
annually and reach 10% of global GDP within the next 10 years (WTTC,
2012). As with all tourism, that which is associated with the Ocean is
heavily influenced by climate change, global economic and socio-political
conditions, and their interactions (Scott et al., 2012b; Section 10.6.1).
Climate change, through impacts on ecosystems (e.g., coral reef bleaching),
can reduce the appeal of destinations, increase operating costs, and/or
increase uncertainty in a highly sensitive business environment (Scott
et al., 2012b).
Several facets of the influence of climate change on the Ocean directly
impact tourism (Section 10.6). Tourism is susceptible to extreme
events such as violent storms, long periods of drought, and/or extreme
precipitation events (Sections 5.4.3.4, 10.6.1; IPCC, 2012). SLR, through
its influence on coastal erosion and submergence, salinization of water
supplies, and changes to storm surge, increases the vulnerability of
coastal tourism infrastructure, tourist safety, and iconic ecosystems
(high confidence; Sections 5.3.3.2, 5.4.3.4, 10.6; Table SPM.1; IPCC,
2012). For example, approximately 29% of resorts in the Caribbean are
within 1 m of the high tide mark and 60% are at risk of beach erosion
from rapid SLR (Scott et al., 2012a).
Increasing sea temperatures (Section 30.3.1.1) can change attractiveness
of locations and the opportunities for tourism through their influence
on the movement of organisms and the state of ecosystems such as
coral reefs (Section 10.6.2; Box CC-CR; UNWTO and UNEP, 2008). Mass
coral bleaching and mortality (triggered by elevated sea temperatures;
high confidence) can decrease the appeal of destinations for diving-
related tourism, although the level of awareness of tourists of impacts
(e.g., <50% of tourists were concerned about coral bleaching during a
major bleaching year, 1998) and expected economic impacts have been
found to be uncertain (Scott et al., 2012b). Some studies, however, have
noted reduced tourist satisfaction and identified “dead coral as one
of the reasons for disappointment at the end of the holiday (Westmacott
et al., 2000). Tourists respond to changes in factors such as weather and
opportunity by expressing different preferences. For example, preferred
conditions and hence tourism are projected to shift toward higher
latitudes with climate change, or from summer to cooler seasons
(Amelung et al., 2007; Section 10.6.1).
Options for adaptation by the marine tourism sector include (1)
identifying and responding to inundation risks with current infrastructure,
and planning for projected SLR when building new tourism infrastructure
(Section 5.5; Scott et al., 2012a); (2) promoting shoreline stability and
natural barriers by preserving ecosystems such as mangroves, salt
marshes, and coral reefs (Section 5.5; Scott et al., 2012b); (3) deploying
forecasting and early-warning systems in order to anticipate challenges
to tourism and natural ecosystems (Strong et al., 2011; IPCC, 2012); (4)
preparation of risk management and disaster preparation plans in order
1705
The Ocean Chapter 30
30
t
o respond to extreme events; (5) reducing the effect of other stressors
on ecosystems and building resilience in iconic tourism features such
as coral reefs and mangroves; and (6) educating tourists to improve
understanding of the negative consequences of climate change over
those stemming from local stresses (Scott et al., 2012a,b). Adaptation
plans for tourism industries need to address specific operators and
regions. For example, some operators may have costly infrastructure at
risk while others may have few assets but are dependent on the
integrity of natural environments and ecosystems (Turton et al., 2010).
30.6.2.3. Shipping
International shipping accounts for more than 80% of world trade by
volume (UNCTAD, 2009a,b) and approximately 3% of global CO
2
emissions from fuel combustion although CO
2
emissions are expected
to increase two- to threefold by 2050 (Heitmann and Khalilian, 2010;
WGIII AR5 Section 8.1). Changes in shipping routes (Borgerson, 2008)
and variation in the transport network due to shifts in grain production
and global markets, as well as new fuel and weather-monitoring
technology, may alter these emission patterns (WGIII AR5 Sections 8.3,
8.5). Extreme weather events, intensified by climate change, may
interrupt ports and transport routes more frequently, damaging
infrastructure and introducing additional dangers to ships, crews, and
the environment (UNCTAD, 2009a,b; Pinnegar et al., 2012; Section
10.4.4). These issues have been assessed by some countries which have
raised concerns over the potential for costly delays and cancellation of
services, and the implications for insurance premiums as storminess and
other factors increase risks (Thornes et al., 2012).
Climate change may benefit maritime transport by reducing Arctic sea ice
and consequently shorten travel distances between key ports (Borgerson,
2008), thus also decreasing total GHG emissions from ships (WGIII AR5
Section 8.5.1). Currently, the low level of reliability of this route limits
its use (Schøyen and Bråthen, 2011), and the potential full operation of
the Northwest Passage and Northern Sea Route would require a transit
management regime, regulation (e.g., navigation, environmental, safety,
and security issues), and a clear legal framework to address potential
territorial claims that may arise, with a number of countries having
direct interest in the Arctic. Further discussion of issues around melting
Arctic sea ice and the Northern Sea Route are given in Chapter 28
(Sections 28.2.6, 28.3.4).
30.6.2.4. Offshore Energy and Mineral Resource
Extraction and Supply
The marine oil and gas industry face potential impacts from climate
change on its ocean-based activities. More than 100 oil and gas
platforms were destroyed in the Gulf of Mexico by the unusually strong
Hurricanes Katrina and Rita in 2005. Other consequences for oil
pipelines and production facilities ultimately reduced US refining
capacity by 20% (IPCC, 2012). The increasing demand for oil and gas
has pushed operations to waters 2000 m deep or more, far beyond
continental shelves. The very large-scale moored developments required
are exposed to greater hazards and higher risks, most of which are not
well understood by existing climate/weather projections. Although there
i
s a strong trend toward seafloor well completions with a complex of
wells, manifolds, and pipes that are not exposed to surface forcing,
these systems face different hazards from instability and scouring of
the unconsolidated sediments by DS currents (Randolph et al., 2010).
The influence of warming oceans on sea floor stability is widely debated
due largely to uncertainties about the effects of methane and methane
hydrates (Sultan et al., 2004; Archer et al., 2009; Geresi et al., 2009).
Declining sea ice is also opening up the Arctic to further oil and gas
extraction. Discussion of potential expansion of oil and mineral production
in the Arctic is made in Chapter 28 (Sections 28.2.5-6, 28.3.4).
The principal threat to oil and gas extraction and infrastructure in
maritime settings is the impact of extreme weather (Kessler et al., 2011),
which is likely to increase given that future storm systems are expected
to have greater energy (Emanuel, 2005; Trenberth and Shea, 2006;
Knutson et al., 2010). Events such as Hurricane Katrina have illustrated
challenges which will arise for this industry with projected increases in
storm intensity (Cruz and Krausmann, 2008). In this regard, early
warning systems and integrated planning offer some potential to reduce
the effect of extreme events (IPCC, 2012).
30.6.3. Human Health
Major threats to public health due to climate change include diminished
security of water and food supplies, extreme weather events, and
changes in the distribution and severity of diseases, including those due
to marine biotoxins (Costello et al., 2009; Sections 5.4.3.5, 6.4.2.3, 11.2).
The predominantly negative impacts of disease for human communities
are expected to be more serious in low-income areas such as Southeast
Asia, southern and east Africa, and various sub-regions of South America
(Patz et al., 2005), which also have under-resourced health systems
(Costello et al., 2009). Many of the influences are directly or indirectly
related to basin-scale changes in the Ocean (e.g., temperature, rainfall,
plankton populations, SLR, and ocean circulation; McMichael et al.,
2006). Climate change in the Ocean may influence the distribution of
diseases such as cholera (Section 11.5.2.1), and the distribution and
occurrence of HABs. The frequency of cholera outbreaks induced by
Vibrio cholerae and other enteric pathogens are correlated with sea
surface temperatures, multi-decadal fluctuations of ENSO, and plankton
blooms, which may provide insight into how this disease may change
with projected rates of ocean warming (Colwell, 1996; Pascual et al.,
2000; Rodó et al., 2002; Patz et al., 2005; Myers and Patz, 2009; Baker-
Austin et al., 2012). The incidence of diseases such as ciguatera also
shows links to ENSO, with ciguatera becoming more prominent after
periods of elevated sea temperature. This indicates that ciguatera may
become more frequent in a warmer climate (Llewellyn, 2010), particularly
given the higher prevalence of ciguatera in areas with degraded coral
reefs (low confidence; Pratchett et al., 2011a).
30.6.4. Ocean-Based Mitigation
30.6.4.1. Deep Sea Carbon Sequestration
Carbon dioxide capture and storage into the deep sea and geologic
structures are also discussed in WGIII AR5 Chapter 7 (Sections 7.5.5, 7.8.2,
1706
Chapter 30 The Ocean
30
7
.12). The economic impact of deliberate CO
2
s
equestration beneath
the sea floor has previously been reviewed (IPCC, 2005). Active CO
2
sequestration from co-produced CO
2
into sub-sea geologic formations is
being instigated in the North Sea and in the Santos Basin offshore from
Brazil. These activities will increase as offshore oil and gas production
increasingly exploits fields with high CO
2
in the source gas and oil.
Significant risks from the injection of high levels of CO
2
into deep ocean
waters have been identified for DS organisms and ecosystems although
chronic effects have not yet been studied. These risks are similar to those
discussed previously with respect to ocean acidification and could
further exacerbate declining O
2
levels and changing trophic networks
in deep water areas (Seibel and Walsh, 2001; Section 6.4.2.2).
There are significant issues within the decision frameworks regulating
these activities. Dumping of any waste or other matter in the sea,
including the seabed and its subsoil, is strictly prohibited under the 1996
London Protocol (LP) except for those few materials listed in Annex I.
Annex 1 was amended in 2006 to permit storage of CO
2
under the
seabed. “Specific Guidelines for Assessment of Carbon Dioxide Streams
for Disposal into Sub-Seabed Geological Formations” were adopted by
the parties to the LP in 2007. The Guidelines take a precautionary
approach to the process, requiring Contracting Parties under whose
jurisdiction or control such activities are conducted to issue a permit
for the disposal subject to stringent conditions being fulfilled (Rayfuse
and Warner, 2012).
30.6.4.2. Offshore Renewable Energy
Renewable energy supply from the Ocean includes ocean energy and
offshore wind turbines. The global technical potential for ocean and
wind energy is not as high as solar energy although considerable
potential still remains. Detailed discussion of the potential of renewable
energy sources are given in WGIII AR5 Chapter 7 (Sections 7.4.2, 7.5.3,
7.8.2). There is an increasing trend in the renewable energy sector to
offshore wind turbines (Section 10.2.2). At present, there is high
uncertainty about how changes in wind intensity and patterns, and
extreme events (from climate change), will impact the offshore wind
energy sector. Given the design and engineering solutions available to
combat climate change impacts (Tables 10-1, 10-7), it is unlikely that this
sector will face insurmountable challenges from climate change.
30.6.5. Maritime Security and Related Operations
Climate change and its influence on the Ocean has become an area of
increasing concern in terms of the maintenance of national security and
the protection of citizens. These concerns have arisen as nation-states
increasingly engage in operations ranging from humanitarian assistance
in climate-related disasters to territorial issues exacerbated by changing
coastlines, human communities, resource access, and new seaways
(Kaye, 2012; Rahman, 2012; Section 12.6). In this regard, increasing sea
levels along gently sloping coastlines can have the seemingly perverse
outcome that the territorial limits to the maritime jurisdiction of the
State might be open to question as the distance from national baselines
to the outer limits of the EEZ increases beyond 200 nm over time
(Schofield and Arsana, 2012).
C
hanges in coastal resources may also be coupled with decreasing
food security to compound coastal poverty and lead, in some cases, to
increased criminal activities such as piracy; IUU fishing; and human,
arms, and drug trafficking (Kaye, 2012). While the linkages have not
been clearly defined in all cases, it is possible that changes in the Ocean
as result of climate change will increase pressure on resources aimed
at maintaining maritime security and countering criminal activity,
disaster relief operations, and freedom of navigation (Section 12.6.2).
National maritime security capacity and infrastructure may also require
rethinking as new challenges present themselves as a result of climate
change and ocean acidification (Allen and Bergin, 2009; Rahman, 2012;
Sections 12.6.1-2).
Opportunities may also arise from changes to international geography
such as formation of new ice-free seaways through the Arctic, which
may benefit some countries in terms of maintaining maritime security
and access (Section 28.2.6). Conversely, such new features may also lead
to increasing international tensions as States perceive new vulnerabilities
from these changes to geography.
Like commercial shipping (Section 30.6.2.3), naval operations in many
countries result in significant GHG emissions (e.g., the US Navy emits
around 2% of the national GHG emissions; Mabus, 2010). As a result,
there are a number of programs being implemented by navies around
the world to try and reduce their carbon footprint and air pollution such
as improving engine efficiency, reducing fouling of vessels, increasing
the use of biofuels, and using nuclear technology for power generation,
among other initiatives.
30.7. Synthesis and Conclusions
Evidence that human activities are fundamentally changing the Ocean
is virtually certain. Sea temperatures have increased rapidly over the
past 60 years at the same time as pH has declined, consistent with the
expected influence of rising atmospheric concentrations of CO
2
and
other GHGs (very high confidence). The rapid rate at which these
fundamental physical and chemical parameters of the Ocean are
changing is unprecedented within the last 65 Ma (high confidence) and
possibly 300 Ma (medium confidence). As the heat content of the
Ocean has increased, the Ocean has become more stratified (very likely),
although there is considerable regional variability. In some cases,
changing surface wind has influenced the extent of mixing and upwelling,
although our understanding of where and why these differences occur
regionally is uncertain. The changing structure and function of the
Ocean has led to changes in parameters such as O
2
, carbonate ion, and
inorganic nutrient concentrations (high confidence). Not surprisingly,
these fundamental changes have resulted in responses by key marine
organisms, ecosystems, and ecological processes, with negative
implications for hundreds of millions of people that depend on the
ecosystem goods and services provided (very likely). Marine organisms
are migrating at rapid rates toward higher latitudes, fisheries are
transforming, and many organisms are shifting their reproductive and
migratory activity in time and in concert with changes in temperature
and other parameters. Ecosystems such as coral reefs are declining
rapidly (high confidence). An extensive discussion of these changes is
provided in previous sections and in other chapters of AR5.
1707
The Ocean Chapter 30
30
C
OO
C
OO
C
OO
Key risk Adaptation issues & prospects
Climatic
drivers
Risk & potential for
adaptation
Timeframe
Risks to ecosystems and adaptation options
P
resent
2°C
4°C
Very
low
V
ery
high
Medium
Present
2°C
4°C
V
ery
low
Very
h
igh
M
edium
Present
2°C
4°C
Very
low
V
ery
high
M
edium
Present
2°C
4°C
Very
low
Very
high
Medium
Present
2°C
4°C
Very
low
Very
high
Medium
Present
2°C
4°C
Very
low
Very
high
Medium
Table 30-3 | Key risks to ocean and coastal issues from climate change and the potential for risk reduction through mitigation and adaptation. Key risks are identified based on
assessment of the literature and expert judgments made by authors of the various WGII AR5 chapters, with supporting evaluation of evidence and agreement in the referenced
chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here,
for 2030–2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the
longer term era of climate options (here, for 2080–2100), for global mean temperature increases of 2°C and 4°C above pre-industrial levels. For each time frame, risk levels are
estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and
human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols.
Near term
(
2030 – 2040)
Long term
(2080 2100)
N
ear term
(20302040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
C
hanges in ecosystem productivity
associated with the redistribution and
loss of net primary productivity in open
oceans. (medium confidence)
[
6.5.1, 6.3.4, Box CC-PP]
A
daptation options are limited to the translocation of industrial fishing activities
due to regional decreases (low latitude) versus increases (high latitude) in
productivity, or to the expansion of aquaculture.
D
istributional shift in fish and invertebrate
s
pecies, fall in fisheries catch potential at
low latitudes, e.g., in EUS, CBS, and STG
regions. (high confidence)
[
6.3.1, Box CC-MB]
E
volutionary adaptation potential of fish and invertebrate species to warming is limited
as indicated by their changes in distribution to maintain temperatures. Human adaptation
o
ptions involve the large-scale translocation of industrial fishing activities following the
regional decreases (low latitude) versus (possibly transient) increases (high latitude) in
c
atch potential as well as deploying flexible management that can react to variability and
change. Further options include improvingsh resilience to thermal stress by reducing
other stressors such as pollution and eutrophication, the expansion of sustainable
aquaculture and development of alternative livelihoods in some regions.
High mortalities and loss of habitat to
larger fauna including commercial species
due to hypoxia expansion and effects.
(high confidence)
[6.3.3, 30.5.3.2, 30.5.4.1-2]
Human adaptation options involve the large-scale translocation of industrial
fishing activities as a consequence of the hypoxia-induced decreases in
biodiversity and fisheries catch of pelagic fish and squid. Special fisheries may
benefit (Humboldt squid). Reducing the amount of organic carbon running off of
coastlines by controlling nutrients and pollution running off agricultural areas can
reduce microbial activity and consequently limit the extent of the oxygen
drawdown and the formation of coastal dead zones.
Ocean acidification: Reduced growth and
survival of commercially valuable shellfish
and other calcifiers, e.g., reef building
corals, calcareous red algae.
(high confidence)
[5.3.3.5, 6.1.1, 6.3.2, 6.4.1.1, 30.3.2.2,
Box CC-OA]
Evidence for differential resistance and evolutionary adaptation of some species
exists but is likely limited by the CO
2
concentrations and high temperatures
reached; adaptation options shifting to exploit more resilient species or the
protection of habitats with low natural CO
2
levels, as well as the reduction of
other stresses, mainly pollution and limiting pressures from tourism and fishing.
Reduced biodiversity, fisheries abundance
and coastal protection by coral reefs due
to heat-induced mass coral bleaching and
mortality increases, exacerbated by ocean
acidification, e.g., in CBS, SES, and STG
regions. (high confidence)
[5.4.2.4, 6.4.2, 30.3.1.1, 30.3.2.2, 30.5.2,
30.5.3, 30.5.4, 30.5.6, Box CC-CR]
Evidence of rapid evolution by corals is very limited or nonexistent. Some corals
may migrate to higher latitudes. However, the movement of entire reef systems is
unlikely given estimates that they need to move at the speed of 10 – 20 km yr
–1
to keep up with the pace of climate change. Human adaptation options are
limited to reducing other stresses, mainly enhancing water quality and limiting
pressures from tourism and fishing. This option will delay the impacts of climate
change by a few decades but is likely to disappear as thermal stress increases.
Coastal inundation and habitat loss due
to sea level rise, extreme events, changes
in precipitation, and reduced ecological
resilience, e.g., in CBS and STG
subregions. (medium to high confidence)
[5.5.2, 5.5.4, 30.5.6.1.3, 30.6.2.2, Box
CC-CR]
Options to maintain ecosystem integrity are limited to the reduction of other
stresses, mainly pollution and limiting pressures from tourism, fishing, physical
destruction, and unsustainable aquaculture. Reducing deforestation and increasing
reforestation of river catchments and coastal areas to retain sediments and
nutrients. Increased mangrove, coral reef, and seagrass protection and restoration
to protect numerous ecosystem goods and services such as coastal protection,
tourist value, and fish habitat.
D
amaging
c
yclone
O
cean
a
cidification
Precipitation
C
O
O
Climate-related drivers of impacts
W
arming
t
rend
E
xtreme
p
recipitation
E
xtreme
t
emperature
S
ea
l
evel
Level of risk & potential for adaptation
P
otential for additional adaptation
t
o reduce risk
Risk level with
c
urrent adaptation
R
isk level with
high adaptation
Hypoxia
O
2
Continued next page
C
OO
Near term
(20302040)
Long term
(2080 2100)
Present
2°C
4°C
Very
low
Very
high
Medium
Marine biodiversity loss with high rate of
climate change. (medium confidence)
[6.3.1-3, 6.4.1.2-3, Table 30.4, Box
CC-MB]
Adaptation options are limited to the reduction of other stresses, mainly to
reducing pollution and to limiting pressures from tourism and fishing.
O
2
1708
Chapter 30 The Ocean
30
30.7.1. Key Risks and Vulnerabilities
The rapid changes in the physical, chemical, and biological state of the
Ocean pose a number of key risks and vulnerabilities for ecosystems,
communities, and nations worldwide. Table 30-3 and Figure 30-12
summarize risks and vulnerabilities from climate change and ocean
acidification, along with adaptation issues and prospects, and a summary
of expert opinion on how these risks will change under further changes
in environmental conditions.
Rising ocean temperatures are changing the distribution, abundance, and
phenology of many marine species and ecosystems, and consequently
represent a key risk to food resources, coastal livelihoods, and industries
such as tourism and fishing, especially for HLSBS, CBS, STG, and EBUE
(Sections 6.3.1, 6.3.4, 7.3.2.4, 30.5; Figure 30-12; Table 30-3; Box CC-MB).
Key risks involve changes in the distribution and abundance of key
fishery species (high confidence; Section 30.6.2.1; Figure 30-12 A,B,G,H)
as well as the spread of disease and invading organisms, each of which
has the potential to impact ecosystems as well as aquaculture and fishing
(Sections 6.3.5, 6.4.1.1, 6.5.3, 7.3.2.4, 7.4.2, 29.5.3-4; Table 30-3).
Adaptation to these changes may be possible in the short-term through
dynamic fisheries policy and management (i.e., relocation of fishing
effort; Table 30-3), as well as monitoring and responding to potential
invading species in coastal settings. The increasing frequency of thermal
extremes (Box CC-HS) will also increase the risk that the thermal
threshold of corals and other organisms is exceeded on a more frequent
Present
2
°C
4°C
V
ery
low
Very
h
igh
M
edium
Present
2
°C
4°C
V
ery
low
V
ery
high
Medium
C
OO
C
O
O
Redistribution of catch potential of large
pelagic-highly migratory fish resources, such as
t
ropical Pacific tuna fisheries. (high confidence)
[6.3.1, 6.4.3, Table 30.4]
International fisheries agreements and instruments, such as the tuna
commissions, may have limited success in establishing sustainable
sheries yields.
Variability of small pelagic fishes in EBUEs is
becoming more extreme at interannual to
multidecadal scales, making industry and
management decisions more uncertain. (medium
confidence)
[6.3.2, 6.3.3, 30.5.2, 30.5.5, Box CC-UP]
D
evelopment of new and specific management tools and models
may have limited success to sustain yields. Reduction in fishing
intensity increases resilience of the fisheries.
Present
2°C
4°C
Very
low
Very
high
Medium
Present
2°C
4°C
Very
low
Very
high
Medium
Decrease in catch and species diversity of fisheries in
tropical coral reefs, exacerbated by interactions with
other human drivers such as eutrophication and
habitat destruction. (high confidence)
[6.4.1, 30.5.3-4, 30.5.6, Box CC-CR]
Restoration of overexploited fisheries and reduction of other
stressors on coral reefs delay ecosystem changes. Human adaptation
includes the usage of alternative livelihoods and food sources (e.g.,
coastal aquaculture).
Current spatial management units, especially the
marine protected areas (MPAs), may fail in the future
due to shifts in species distributions and community
structure.
(high confidence)
[6.3.1, 6.4.2.1, 30.5.1, Box CC-MB]
Continuous revision and shifts of MPA borders, and of MPA goals
and performance.
Present
2°C
4°C
V
ery
low
Very
h
igh
M
edium
P
resent
2°C
4°C
Very
low
Very
high
Medium
Decreased production of global shellfish fisheries.
(high confidence)
[6.3.2, 6.3.5, 6.4.1.1, 30.5.5, 30.6.2.1, Box CC-OA]
Effective shift to alternative livelihoods, changes in food
consumption patterns, and adjustment of (global) markets.
Global redistribution and decrease of low-latitude
fisheries yields are paralleled by a global trend to
c
atches having smaller fishes. (medium confidence)
[6.3.1, 6.4.1, 6.5.3, 30.5.4, 30.5.6, 30.6.2]
Increasing coastal poverty at low latitudes as fisheries becomes
smaller – partially compensated by the growth of aquaculture and
m
arine spatial planning, as well as enhanced industrialized fishing
e
fforts.
C
OO
C
OO
O
2
O
2
N
ear term
(20302040)
Long term
(
2080 2100)
Near term
(
2030 – 2040)
L
ong term
(2080 2100)
Near term
(20302040)
L
ong term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Key risk Adaptation issues & prospects
Climatic
drivers
Risk & potential for
adaptation
Timeframe
Risks to sheries
Continued next page
Table 30-3 (continued)
Continued next page
1709
The Ocean Chapter 30
30
basis (especially in CBS, STG, SES, HLSBS, and EUS regions; Sections 6.2,
30.5; Box CC-CR). These changes pose a key risk to vulnerable ecosystems
such as mangroves and coral reefs, with potential to have a series of
serious impacts on fisheries, tourism, and coastal ecosystem services
such as coastal protection (Sections 5.4.2.4, 6.3.2, 6.3.5, 6.4.1.3, 7.2.1.2,
29.3.1.2, 30.5; Table 30-3; Box CC-CR). Genetic adaptation of species
to increasing levels of stress may not occur fast enough given fairly long
generation times of organisms such as reef-building corals and many
other invertebrates and fish (Table 30-3). In this case, risks may be
reduced by addressing stresses not related to climate change (e.g.,
pollution, overfishing), although this strategy could have minimal
impact if further increases in sea temperature occur (high confidence).
Loss of these important coastal ecosystems is associated with emerging
risks associated with the collapse of some coastal fisheries along with
livelihoods, food, and regional security (medium confidence). These
changes are likely to be exacerbated by other key risks such as coastal
inundation and habitat loss due to SLR, as well as intensified precipitation
events (high confidence; Section 5.4; Box CC-CR). Adaptation options
in this case include engineered coastal defenses, reestablishing coastal
vegetation such as mangroves, protecting water supplies from salination,
and developing strategies for coastal communities to withdraw to less
vulnerable locations over time (Section 5.5).
The recent decline in O
2
concentrations has been ascribed to warming
through the effect on ocean mixing and ventilation, as well as the
solubility of O
2
and its consumption by marine microbes (Sections 6.1.1.3,
6.3.3, 30.3.2.3, 30.5.7). This represents a key risk to ocean ecosystems
(medium confidence; Figure 30-12 5,6,C). These changes increase the
vulnerability of marine communities, especially those below the
euphotic zone, to hypoxia and ultimately lead to a restriction of suitable
habitat (high confidence; Figure 30-12 5). In the more extreme case,
often exacerbated by the contribution of organic carbon from land-based
sources, “dead zones” may form. Decreasing oxygen, consequently, is
very likely to increase the vulnerability of fisheries and aquaculture
(medium confidence; Figure 30-12 C), and consequently puts livelihoods
Key risk Adaptation issues & prospects
Climatic
drivers
Risk & potential for
adaptation
Timeframe
Risks to humans and infrastructure (continued)
C
OO
P
resent
2°C
4°C
Very
low
Very
h
igh
Medium
Present
2°C
4°C
Very
low
V
ery
high
M
edium
Impacts due to increased frequency of
harmful algal blooms (medium confidence)
[6.4.2.3]
Adaptation options include improved monitoring and early warning system,
reduction of stresses favoring harmful algal blooms, mainly pollution and
eutrophication, as well as the avoidance of contaminated areas and fisheries
products.
Impacts on marine resources threatening
regional security as territorial disputes and
food security challenges increase
(limited evidence, medium agreement)
[IPCC 2012, 30.6.5, 12.4-12.6, 29.3]
Decrease in marine resources, movements of fish stocks and opening of new
seaways , and impacts of extreme events coupled with increasing populations
will increase the potential for conflict in some regions, drive potential
migration of people, and increase humanitarian crises.
C
OO
Present
2°C
4°C
Very
low
Very
high
Medium
Impacts on shipping and infrastructure for
energy and mineral extraction increases as
storm intensity and wave height increase
in some regions (e.g., high latitudes)
(high confidence)
[IPCC 2012, 30.6.5, 12.4-12.6, 29.3]
Adaptation options are to limit activities to particular times of the year and/or
develop strategies to decrease the vulnerability of structures and operations.
Near term
(20302040)
Long term
(2080 2100)
Near term
(
2030 – 2040)
Long term
(2080 2100)
Near term
(20302040)
Long term
(2080 2100)
Table 30-3 (continued)
P
resent
2°C
4°C
V
ery
low
Very
high
M
edium
Reduced livelihoods and increased poverty.
(
medium confidence)
[6.4.1-2, 30.6.2, 30.6.5]
Human adaptation options involve the large-scale translocation of industrial
shing activities following the regional decreases (low latitude) versus increases
(
high latitude) in catch potential and shifts in biodiversity. Artisanal fisheries are
extremely limited in their adaptation options by available financial resources
and technical capacities, except for their potential shift to other species of
interest.
N
ear term
(20302040)
Long term
(
2080 2100)
P
resent
2°C
4°C
Very
low
V
ery
high
Medium
Reduced coastal socioeconomic security.
(
high confidence)
[5.5.2, 5.5.4, 30.6.5, 30.7.1]
Human adaptation options involve (1) protection using coastal defences (e.g. seawalls
w
here appropriate and economic) and soft measures (e.g., mangrove replanting and
enhancing coral growth); (2) accommodation to allow continued occupation of
c
oastal areas by making changes to human activities and infrastructure; and (3)
m
anaged retreat as a last viable option. Vary from large-scale engineering works to
smaller scale community projects. Options are available under the more traditional
C
ZM (coastal zone management) framework but increasingly under DRR (disaster risk
reduction) and CCA (climate change adaptation) frameworks.
*H
igh confidence in existence of adaptation measures, Low confidence in magnitude of risk reduction
*
*
*
*
Near term
(
2030 – 2040)
L
ong term
(
2080 2100)
CBS = Coastal Boundary Systems; EBUE = Eastern Boundary Upwehlling Ecosystems; EUS = Equatorial Upwelling Systems; HLSBS = High-Latitude Spring Bloom Systems;
SES = Semi-Enclosed Seas; STG = Subtropical Gyres.
1710
Chapter 30 The Ocean
30
a
t risk, particularly in EBUE (e.g., California and Humboldt Current
ecosystems; Section 30.5.5), SES (e.g., Baltic and Black Seas; Section
30.5.3), and CBS (e.g., Gulf of Mexico, northeast Indian Ocean; Sections
30.3.2.3, 30.5.4). It is very likely that the warming of surface waters has
also increased the stratification of the upper ocean by about 4% between
0 and 200 m from 1971 to 2010 in all oceans north of about 40°S. In
many cases, there is significant adaptation opportunity to reduce
hypoxia locally by reducing the flow of organic carbon, hence microbial
activity, within these coastal systems (Section 30.5.4). Relocating fishing
effort, and modifying procedures associated with industries such as
aquaculture, may offer some opportunity to adapt to these changes
(likely). Declining O
2
concentrations are likely to have significant impacts
on DS habitats, where organisms are relatively sensitive to environmental
changes of this nature owing to the very constant conditions under
which they have evolved (Section 30.5.7).
Ocean acidification has increased the vulnerability of ocean ecosystems
by affecting key aspects of the physiology and ecology of marine
organisms (particularly in CBS, STG, and SES; Section 6.3.2; Table 30-3;
Box CC-OA). Decreasing pH and carbonate ion concentrations reduce
the ability of marine organisms to produce shells and skeletons, and
may interfere with a range of biological processes such as reproduction,
gas exchange, metabolism, navigation ability, and neural function in a
broad range of marine organisms that show minor to major influences of
ocean acidification on their biology (Sections 6.3.2, 30.3.2.2; Box CC-OA).
Natural variability in ocean pH can interact with ocean acidification to
create damaging periods of extremes (i.e., high CO
2
, low O
2
and pH),
which can have a strong effect on coastal activities such as aquaculture
(medium confidence; Section 6.2; Figure 30-12 A; Box CC-UP). There may
be opportunity to adapt aquaculture to increasingly acidic conditions
by monitoring natural variability and restricting water intake to periods
of optimal conditions. Reducing other non-climate change or ocean
acidification associated stresses also represents an opportunity to build
greater ecological resilience against the impacts of changing ocean
carbonate chemistry. Ocean acidification is also an emerging risk for
DS habitats as CO
2
continues to penetrate the Ocean, although the
impacts and adaptation options are poorly understood and explored.
Ocean acidification has heightened importance for some groups of
organisms and ecosystems (Box CC-OA). In ecosystems that are heavily
dependent on the accumulation of calcium carbonate over time (e.g.,
c
oral reefs, Halimeda beds), increasing ocean acidification puts at risk
ecosystems services that are critical for hundreds of thousands of
marine species, plus people and industries, particularly within CBS, STG,
and SES (high confidence). Further risks may emerge from the non-
linear interaction of different factors (e.g., increasing ocean temperature
may amplify effects of ocean acidification, and vice versa) and via the
interaction of local stressors with climate change (e.g., interacting
changes may lead to greater ecosystems disturbances than each impact
on its own). There is an urgent need to understand these types of
interactions and impacts, especially given the long time it will take to
return ocean ecosystems to preindustrial pH and carbonate chemistry
(i.e., tens of thousands of years (FAQ 30.1) should CO
2
emissions continue
at the current rate).
It is very likely that surface warming has increased stratification of
the upper ocean, contributing to the decrease in O
2
along with the
temperature-related decreases in oxygen solubility (WGI AR5 Section
3.8.3). Changes to wind speed, wave height, and storm intensity influence
the location and rate of mixing within the upper layers of the Ocean
and hence the concentration of inorganic nutrients (e.g., in EBUE, EUS;
Figure 30-12 1,3). These changes to ocean structure increase the risks and
vulnerability of food webs within the Ocean. However, our understanding
of how primary productivity is going to change in a warming and more
acidified ocean is limited, as is our understanding of how upwelling will
respond to changing surface wind as the world continues to warm
(Boxes CC-PP, CC-UP). As already discussed, these types of changes can
have implications for the supply of O
2
into the Ocean and the upward
transport of inorganic nutrients to the euphotic zone. Although our
understanding is limited, there is significant potential for regional increases
in wind speed to result in greater rates of upwelling and the supply of
inorganic nutrients to the photic zone. Although this may increase
productivity of phytoplankton communities and associated fisheries,
greater rates of upwelling can increase the risk of hypoxic conditions
developing at depth as excess primary production sinks into the Ocean
and stimulates microbial activity at depth (Sections 6.1.1.3, 30.3.2.3,
30.5.5; Table 30-3). Changes in storm intensity may increase the risk of
damage to shipping and industrial infrastructure, which increases the risk
of accidents and delays to the transport of products between countries,
security operations, and the extraction of minerals from coastal and
oceanic areas (Section 30.6.2; IPCC, 2012).
Frequently Asked Questions
FAQ 30.5 | How can we use non-climate factors to manage
climate change impacts on the oceans?
The Ocean is exposed to a range of stresses that may or may not be related to climate change. Human activities
can result in pollution, eutrophication (too many nutrients), habitat destruction, invasive species, destructive fishing,
and over-exploitation of marine resources. Sometimes, these activities can increase the impacts of climate change,
although they can, in a few circumstances, dampen the effects as well. Understanding how these factors interact
with climate change and ocean acidification is important in its own right. However, reducing the impact of these
non-climate factors may reduce the overall rate of change within ocean ecosystems. Building ecological resilience
through ecosystem-based approaches to the management of the marine environment, for example, may pay dividends
in terms of reducing and delaying the effects of climate change (high confidence).
1711
The Ocean Chapter 30
30
T
he proliferation of key risks and vulnerabilities to the goods and
services provided by ocean ecosystems as a result of ocean warming and
acidification generate a number of key risks for the citizens of almost
every nation. Risks to food security and livelihoods are expected to
increase over time, aggravating poverty and inequity (Table 30-3). As
these problems increase, regional security is likely to deteriorate as
disputes over resources increase, along with increasing insecurity of
food and nutrition (Sections 12.4-6, 29.3.3, 30.6.5; Table 30-3; IPCC,
2012).
30.7.2. Global Frameworks for Decision Making
Global frameworks for decision making are central to management of
vulnerability and risk at the scale and complexity of the world’s oceans.
General frameworks and conventions for policy development and decision
m
aking within oceanic and coastal regions are important in terms of
the management of stressors not directly due to ocean warming or
acidification, but that may influence the outcome of these two factors.
Tables 30-3 and 30-4 outline a further set of challenges arising from
multiple interacting stressors, as well as potential risks and vulnerabilities,
ramifications, and adaptation options. In the latter case, examples of
potential global frameworks and initiatives for beginning and managing
these adaptation options are described. These frameworks represent
opportunities for global cooperation and the development of international,
regional, and national policy responses to the challenges posed by the
changing ocean (Kenchington and Warner, 2012; Tsamenyi and Hanich,
2012; Warner and Schofield, 2012).
The United Nations Convention on the Law of the Sea (UNCLOS) was a
major outcome of the third UN Conference on the Law of the Sea
(UNCLOS III). The European Union and 164 countries have joined in the
Continued next page
Primary
driver(s)
Biophysical
change
projected
Key risks and
vulnerabilities
Ramifi cations Adaptation options
Policy frameworks and
initiatives (examples)
Key references
and chapter
sections
hT, hOA
Spatial and temporal
v
ariation in primary
productivity (medium
confi dence at global
scales; Box CC-PP)
Reduced fi sheries production
i
mpacts important sources
of income to some countries
while others may see
increased productivity
(e.g., as tuna stocks shift
eastwards in the Pacifi c)
(medium confi dence).
Reduced national
i
ncome, increased
unemployment, plus
increase in poverty.
Potential increase in
disputes over national
ownership of key fi shery
resources (likely)
Increased international cooperation
o
ver key fi sheries. Improved
understanding of linkages between
ocean productivity, recruitment, and
sheries stock levels. Implementation
of the regional “vessel day scheme”
provides social and economic
incentives to fi sheries and fi shers for
adaptation.
UNCLOS, PEMSEA, CTI,
R
FMO agreements, UNSFSA
Bell et al. (2011,
2
013a); Tsamenyi
and Hanich
(2012); Sections
6.4.1, 6.5.3,
30.6.2.1, 30.7.2;
Box CC-PP
hT, hOA
Ecosystem regime
shifts (e.g., coral
to algal reefs;
structural shifts
in phytoplankton
communities)
(medium confi dence)
Reduced fi sheries production
of coastal habitats and
ecosystems such as coral
reefs (medium confi dence).
Decreased food and
employment security and
human migration away
from coastal zone (likely)
Strengthen coastal zone management
to reduce contributing stressors (e.g.,
coastal pollution, over-harvesting, and
physical damage to coastal resources).
Promote Blue Carbon
a
initiatives.
PEMSEA, CTI, PACC,
MARPOL, UNHCR, CBD,
International Organization
for Migration, Global
Environment Facility,
International Labor
Organization
Bell et al. (2013a);
Sections 5.4.3,
6.3.1–2, 12.4,
29.3.1, 29.3.3,
30.5.2–4, 30.5.6,
30.6.1, 30.6.2.1;
Box CC-CR
Tourist appeal of coastal
assets decreases as
ecosystems change to less
“desirable” state, reducing
income to some countries
(low confi dence).
Increased levels of
coastal poverty in some
countries as tourist
income decreases (likely)
As above, strengthen coastal zone
management and reduce additional
stressors on tourist sites; implement
education programs and awareness
among visitors. Diversify tourism
activities.
CBD, PEMSEA, CTI, PACC,
UNHCR, MARPOL
Kenchington and
Warner (2012);
Sections 5.5.4.1,
6.4.1–2, 10.6,
30.6.2.2
Increased risk of some
diseases (e.g., ciguatera,
harmful algal blooms) as
temperatures increase shift
and ecosystems shift away
from coral dominance (low
confi dence).
Increased disease and
mortality; decreases in
coastal food resources
and fi sheries income
(likely)
Increase monitoring and education
surrounding key risks (e.g., ciguatera);
develop alternate fi sheries and income
for periods when disease incidence
increases, and develop or update
health response plans.
National policy strategies
and regional cooperation
needed
Llewellyn (2010);
Sections 6.4.2.3,
10.6, 29.3.3.2,
29.5.3, 30.6.3
Increased poverty and
dislocation of coastal people
(particularly in the tropics)
as coastal resources such as
sheries degrade (medium
confi dence)
Increased population
pressure on migration
destinations (e.g., large
regional cities), and
reduced freedom to
navigate in some areas
(as criminal activity
increases) (likely)
Develop alternative industries and
income for affected coastal people.
Strengthen coastal security both
nationally and across regions. Increase
cooperation over handling of criminal
activities.
UNCLOS, PEMSEA, CTI,
International Ship and Port
Facility Security, IMO, Bali
Process, Association of
Southeast Asian Nations
MLA Treaty and bilateral
extradition and MLA
agreements
Kaye (2012);
Rahman
(2012); Sections
12.4–6,29.3.3,
29.6.2, 30.6.5
Table 30-4 | Ramifi cations, adaptation options, and frameworks for decision making for ocean regions. Symbols for primary drivers: IC = ice cover; NU = nutrient concentration;
OA = ocean acidifi cation; SLR = sea level rise; SS = storm strength; T = sea temperature (h = increased; i = decreased; * = uncertain).
a
Blue Carbon initiatives include conservation and restoration of mangroves, saltmarsh, and seagrass beds as carbon sinks (Section 30.6.1).
Notes: CBD = Convention on Biological Diversity; CTI = Coral Triangle Initiative; IHO = International Hydrographic Organization; IOM = International Organization of Migration;
ISPS = International Ship and Port Facility Security; MARPOL = International Convention for the Prevention of Pollution From Ships; MLA = mutual legal assistance; PACC
= Pacifi cAdaptation to Climate Change Project; PEMSEA = Partnerships in Environmental Management for the Seas of East Asia; RFMO = Regional Fisheries Management
Organizations; UNCLOS = United Nations Convention on the Law of the Sea; UNHCR = United Nations High Commissioner for Refugees; UNSFSA = United Nations Straddling
Fish Stocks Agreement.
1712
Chapter 30 The Ocean
30
Primary
driver(s)
Biophysical
change
projected
Key risks and
vulnerabilities
Ramifi cations Adaptation options
Policy frameworks and
initiatives (examples)
Key references
and chapter
sections
hT
Migration of
o
rganisms and
ecosystems to higher
l
atitudes (high
c
onfi dence)
Reorganization of
c
ommercial fi sh stocks and
ecological regime shifts
(
medium to high confi dence)
Social and economic
d
isruption (very likely)
Increase international cooperation
a
nd improve understanding of
regime changes; implement early-
d
etection monitoring of physical and
b
iological variables and regional
seasonal forecasting; include
related uncertainties into fi sheries
m
anagement; provide social and
economic incentives for industry.
UNCLOS, CBD, RFMO
a
greements, UNSFSA
Sections 7.4.2, 6.5,
3
0.5, 30.6.2.1; Box
CC-MB
I
ncrease in abundance,
g
rowing season, and
distributional extent of pests
a
nd fouling species (medium
confi dence)
I
ncreased disease risk
t
o aquaculture and
sheries. Income loss
a
nd increased operating
and maintenance costs
(
very likely)
I
ncrease environmental monitoring;
p
romote technological advances to
deal with pest and fouling organisms;
i
ncrease vigilance and control related
to biosecurity.
I
MO, ballast water
m
anagement, Anti- Fouling
Convention
S
ections 6.4.1.5,
7
.3.2.4, 29.5.3–4,
30.6.2.1; Box
C
C-MB
Threats to human health
increase due to expansion
o
f pathogen distribution
to higher latitudes (low
c
onfi dence)
Increased disease and
mortality in some coastal
c
ommunities (likely)
Reduce exposure through increased
monitoring and education, adoption,
o
r update of health response plans to
outbreaks.
UNICEF, World Health
Ogranization, IHOs, and
n
ational governments
Myers and Patz
(2009); Sections
6
.4.3, 10.8.2, 11.7,
29.3.3, 30.6.3; Box
C
C-MB
hT, hNU,
hOA*
Increased incidence
o
f harmful algal
blooms (low
c
onfi dence)
Increased threats to
e
cosystems, fi sheries, and
human health (medium
c
onfi dence)
Reduced supply of
m
arine fi sh and shellfi sh
and greater incidence
o
f disease among some
coastal communities
(
likely)
Provide early-detection monitoring
a
nd improve predictive models;
provide education and adoption or
u
pdate of health response plans.
CTI, PEMSEA, World Health
O
rganization, MARPOL
Llewellyn (2010);,
S
ections 30.6.3,
11.7, 6.4.2.3
hT
Increased
precipitation as a
result of intensifi ed
hydrological cycle in
some coastal areas
(medium confi dence)
Increased freshwater,
sediment, and nutrient fl ow
into coastal areas; increase
in number and severity of
ood events (medium to
high confi dence)
Increasing damage to
coastal reef systems with
ecological regime shifts
in many cases (very
likely)
Improve management of catchment
and coastal processes; expand riparian
vegetation along creeks and rivers;
improve agricultural retention of soils
and nutrients.
CTI, PEMSEA, Secretariat
of the Pacifi cRegional
Environment Programme
Sections 3.4,
29.3.1, 30.5.4,
30.6.1
hT
Changing weather
patterns, storm
frequency (medium
confi dence)
Increased risk of damage to
infrastructure such as that
involved in shipping and
oil and gas exploration and
extraction (medium to low
confi dence)
Increased damage and
associated costs (likely)
Adjust infrastructure specifi cations,
develop early-warning systems, and
update emergency response plans to
extreme events.
IMO IPCC (2012);
Sections 10.4.4,
29.3, 30.6.2.3–4
hSLR,
hSS
Increased wave
exposure of coastal
areas and increased
sea level (high
confi dence)
Exposure of coastal
infrastructure and
communities to damage
and inundation, increased
coastal erosion (high
confi dence)
Increased costs to
human towns and
settlements, numbers of
displaced people, and
human migration (very
likely)
Develop integrated coastal
management that considers SLR
in planning and decision making;
increase understanding of the issues
through education.
UNICEF, IHOs, and national
governments
Warner (2012);
Sections 5.5,
12.4.1, 29.5.1,
30.3.1.2, 30.6.5
Inundation of coastal
aquifers reduces water
supplies and decreases
coastal agricultural
productivity (high
confi dence).
Reduced food and
water security leads
to increased coastal
poverty, reduced food
security, and migration
(very likely).
Assist communities in fi nding
alternatives for food and water, or
assist in relocation of populations and
agriculture from vulnerable areas.
UNICEF, IHOs, and national
governments.
Warner (2012);
Sections 5.4.3,
12.4.1, 29.3.2,
30.3.1.2
hSLR
Risk of inundation
and coastal erosion,
especially in low-
lying countries (high
confi dence)
UNCLOS-defi ned limits
of maritime jurisdiction
will contract as national
baselines shift inland.
Potential uncertainty
increases in some areas
with respect to the
international boundaries to
maritime jurisdiction (high
confi dence).
Lack of clarity increases,
as do disputes over
maritime limits and
maritime jurisdiction.
Some nations at risk
of major losses to their
territorial waters (very
likely)
Seek resolution of “shifting national
baselines” issue (retreat and
redefi nition, stabilization, or fi xation
of exclusive economic zones and other
currently defi ned maritime jurisdiction
limits).
UNCLOS IPCC (2012);
Schofi eld and
Arsana (2012);
Warner and
Schofi eld (2012);
Sections 5.5,
30.6.5
hT, iIC
Loss of summer sea
ice (high confi dence)
Access to northern coasts
of Canada, USA, and Russia
increases security concerns
(high confi dence).
Potential for increased
tension on different
interpretations of access
rights and boundaries
(likely to very likely)
Seek early resolution of areas in
dispute currently and in the future.
UNCLOS Chapter 28
New resources become
available as ice retreats,
increasing vulnerability
of international borders
in some cases (medium
confi dence).
Tensions over maritime
claims and ownership of
resources (likely)
Sort out international agreements.
Table 30-4 (continued)
1713
The Ocean Chapter 30
30
C
onvention. UNCLOS replaced earlier frameworks that were built around
the “freedom of the seas” concept and that limited territorial rights to
3 nm off a coastline. UNCLOS provides a comprehensive framework for
the legitimate use of the Ocean and its resources, including maritime
zones, navigational rights, protection and preservation of the marine
environment, fishing activities, marine scientific research, and mineral
resource extraction from the seabed beyond national jurisdiction. The
relationship between climate change and UNCLOS is not clear and
depends on interpretation of key elements within the UNFCCC (United
Nations Framework Convention for Climate Change) and Kyoto Protocol
(Boyle, 2012). However, UNCLOS provides mechanisms to help structural
adaptation in response to challenges posed by climate change. In a similar
way, there is a wide range of other policy and legal frameworks that
structure and enable responses to the outcomes of rapid anthropogenic
climate change in the Ocean.
There are many existing international conventions and agreements that
explicitly recognize climate change (Table 30-4). The UN Straddling Fish
Stocks Agreement (UNSFSA) aims at enhancing international cooperation
of fisheries resources, with an explicit understanding under Article 6 that
management needs to take account “existing and predicted oceanic,
environmental and socio-economic conditions” and to undertake
“relevant research, including surveys of abundance, biomass surveys,
hydro-acoustic surveys, research on environmental factors affecting
stock abundance, and oceanographic and ecological studies” (UNSFSA,
Annex 1, Article 3). International conventions such as these will become
increasingly important as changes to the distribution and abundance
of fisheries are modified by climate change and ocean acidification.
Global frameworks for decision making are increasingly important in
the case of the Ocean, most of which falls outside national boundaries
(Oude Elferink, 2012; Warner, 2012). Approximately 64% of the Ocean
(40% of the Earth’s surface) is outside EEZs and continental shelves of
the world’s nations (high seas and seabed beyond national jurisdiction).
With rapidly increasing levels of exploitation, there are increasing calls
for more effective decision frameworks aimed at regulating fishing and
other activities (e.g., bio-prospecting) within these ocean “commons.”
These international frameworks will become increasingly valuable as
nations respond to impacts on fisheries resources that stretch across
national boundaries. One such example is the multilateral cooperation
that was driven by President Yudhoyono of Indonesia in August 2007
and led to the Coral Triangle Initiative on Coral Reefs, Fisheries, and
Food Security (CTI), which involves region-wide (involving 6.8 million
km
2
including 132,800 km of coastline) cooperation between the
governments of Indonesia, Philippines, Malaysia, Papua New Guinea,
the Solomon Islands, and Timor Leste on reversing the decline in coastal
ecosystems such as coral reefs (Clifton, 2009; Hoegh-Guldberg et al.,
2009; Veron et al., 2009). Partnerships, such as CTI, have the potential
to provide key frameworks to address issues such as interaction between
the over-exploitation of coastal fishing resources and the recovery of
reefs from mass coral bleaching and mortality, and the implications of
the movement of valuable fishery stocks beyond waters under national
jurisdiction.
An initiative called the Global Partnership for Oceans set out to establish
a global framework with which to share experience, resources, and
expertise, as well as to engage governments, industry, civil, and public
s
ector interests in both understanding and finding solutions to key
issues such as overfishing, pollution, and habitat destruction (Hoegh-
Guldberg et al., 2013). Similarly, the Areas Beyond National Jurisdiction
(ABNJ, Global Environment Facility) Initiative has been established to
promote the efficient, collaborative, and sustainable management of
fisheries resources and biodiversity conservation across the Ocean.
Global partnerships are also essential for providing support to the many
nations that often do not have the scientific or financial resources to
solve the challenges that lie ahead (Busby, 2009; Mertz et al., 2009). In
this regard, international networks and partnerships are particularly
significant in terms of assisting nations in developing local adaptation
solutions to their ocean resources. By sharing common experiences and
strategies through global networks, nations have the chance to tap into
a vast array of options with respect to responding to the negative
consequences of climate change and ocean acidification on the world’s
ocean and coastal resources.
30.7.3. Emerging Issues, Data Gaps, and Research Needs
Although there has been an increase in the number of studies being
undertaken to understand the physical, chemical, and biological changes
within the Ocean in response to climate change and ocean acidification,
the number of marine studies of ecological impacts and risks still lag
behind terrestrial studies (Hoegh-Guldberg and Bruno, 2010; Poloczanska
et al., 2013). Rectifying this gap should be a major international objective
given the importance of the Ocean in terms of understanding and
responding to future changes and consequences of ocean warming and
acidification.
30.7.3.1. Changing Variability and Marine Impacts
Understanding the long-term variability of the Ocean is critically important
in terms of the detection and attribution of changes to climate change
(Sections 30.3, 30.5.8), but also in terms of the interaction between
variability and anthropogenic climate change. Developing instrument
systems that expand the spatial and temporal coverage of the Ocean
and key processes will be critical to documenting and understanding
its behavior under further increases in average global temperature
and changes in the atmospheric concentration of CO
2
. International
collaborations such as the Argo network of oceanographic floats
illustrate how international cooperation can rapidly improve our
understanding of the physical behavior of the Ocean and will provide
important insight into its long-term subsurface variability (Schofield et
al., 2013).
30.7.3.2. Surface Wind, Storms, and Upwelling
Improving our understanding of the potential behavior of surface wind
in a warming world is centrally important to our understanding of how
upwelling will change in key regions (e.g., EUS, EBUE; Box CC-UP).
Understanding these changes will provide important information for
future fisheries management but will also illuminate the potential risks
of intensified upwelling leading to hypoxia at depth and the potential
1714
Chapter 30 The Ocean
30
e
xpansion of “dead zones” (Sections 30.3.2, 30.5.2-4). Understanding
surface wind in a warming climate will also yield important information
on surface mixing as well as how surface wave height might also vary,
improving our understanding of potential interactions in coastal areas
between wind, waves, and SLR (Section 30.3.1). Given the importance
of mixing and upwelling to the supply of inorganic nutrients to the
surface layers of the ocean, understanding these important phenomena
at the ocean-atmosphere interface will provide important insight into
how ocean warming and acidification are likely to impact ecosystems,
food webs, and ultimately important fisheries such as those found along
the west coasts of Africa and the Americas.
30.7.3.3. Declining Oxygen Concentrations
The declining level of O
2
in the Ocean is an emerging issue of major
importance (Section 30.3.2). Developing a better understanding of the
role and temperature sensitivity of microbial systems in determining O
2
concentrations will enable a more coherent understanding of the changes
and potential risks to marine ecosystems. Given the importance of
microbial systems to the physical, chemical, and biological characteristics
of the Ocean, it is extremely important that these systems receive
greater focus, especially with regard to their response to ocean warming
and acidification. This is particularly important for the DS (>1000 m),
which is the most extensive habitat on the planet. In this respect,
increasing our understanding of DS habitats and how they may be
changing under the influence of climate change and ocean acidification
is of great importance. Linkages between changes occurring in the surface
layers and those associated with the DS are particularly important in
light of our need to understand how rapidly changes are occurring and
what the implications are for the metabolic activity and O
2
content of
DS habitats.
30.7.3.4. Ocean Acidification
The rapid and largely unprecedented changes to ocean acidification
represent an emerging issue given the central importance of pH and
the concentration of ions such as carbonate in the biology of marine
organisms (Box CC-OA). Despite the relatively short history of research
on this issue, there are already a large number of laboratory and field
studies that demonstrate a large range of effects across organisms,
processes, and ecosystems. Key gaps (Gattuso et al., 2011) remain in
our understanding of how ocean acidification will interact with other
changes in the Ocean, and whether or not biological responses to ocean
acidification are necessarily linear. The vulnerability of fishery species
(e.g., molluscs) to ocean acidification represents an emerging issue, with
a need for research to understand and develop strategies for fishery
and aquaculture industries to minimize the impacts. Understanding of
how carbonate structures such as coral reefs and Halimeda beds will
respond to a rapidly acidifying ocean represents a key gap and research
need, especially in understanding the rate at which consolidated
carbonate structures and related habitats are likely to erode and dissolve.
Interactions between ocean acidification, upwelling, and decreasing O
2
represent additional areas of concern and research. There is also a need
to improve our understanding of the socioeconomic ramifications of
ocean acidification (Turley and Boot, 2011; Hilmi et al., 2013).
30.7.3.5. Net Primary Productivity
Oceanic phytoplankton are responsible for approximately 50% of global
net primary productivity. However, our understanding of how oceanic
primary production is likely to change in a warmer and more acidified
ocean is uncertain (Boxes CC-PP, CC-UP). Changes in net primary
productivity will resonate through food webs and ultimately affect
fisheries production. Given the central role that primary producers and
their associated ecological processes play in ocean ecosystem functioning,
the understanding of how net primary productivity is likely to vary at
global and regional levels is improved (Sections 30.5.2, 30.5.5). At the
same time, understanding how plankton communities will vary spatially
and temporarily will be important in any attempt to understand how fish
populations will fare in a warmer and more acidified ocean. The research
challenge is to determine when and where net primary production is
expected to change, coupled with research on adaptation strategies for
coping with the changes to the global distribution of seafood procurement,
management, and food security.
30.7.3.6. Movement of Marine Organisms and Ecosystems
Marine organisms are moving generally toward higher latitudes or
deeper waters consistent with the expectation of a warming ocean. Our
current understanding of which organisms and ecosystems are moving,
ramifications for reorganization of ecosystems and communities, and
the implications for nations is uncertain at best. Given the implications
for fisheries, invasive species, and the spread of disease, it is imperative
that our understanding of the movement of ecosystems is improved.
Documentation of species’ responses and a deeper understanding of
the processes that lead to persistent range shifts, and a focus on
the ecosystem, social, and economic implications of range shifts is an
important research need.
30.7.3.7. Understanding Cumulative and Synergistic Impacts
Understanding cumulative and synergistic impacts is poorly developed
for ocean systems. Much of our understanding has been built on
experimental approaches that are focused on single stressors that
respond gradually without interaction or impacts that accumulate over
time (Table 30-3). Multifactorial experiments exploring the impact of
combined variables (e.g., elevated temperature and acidification at the
same time) will enable more realistic projections of the future to be
established. Equally, developing a better understanding of how biological
and ecological responses change in relation to key environmental
variables should also be a goal of future research. In this regard,
assumptions that responses are likely to be gradual and linear over
time ultimately have little basis, yet are widespread within the scientific
literature.
30.7.3.8. Reorganization of Ecosystems and Food Webs
The pervasive influence of ocean warming and acidification on the
distribution, abundance, and function of organisms and processes has
and will continue to drive the reorganization of ecosystems and food
1715
The Ocean Chapter 30
30
w
ebs (virtually certain; Hoegh-Guldberg and Bruno, 2010; Poloczanska
et al., 2013; Box CC-MB). One of the inevitable outcomes of differing
tolerances and responses to climate change and ocean acidification is the
development of novel assemblages of organisms in the near future. Such
communities are likely to have no past or contemporary counterparts,
and will consequently require new strategies for managing coastal
areas and fisheries. Changes to a wide array of factors related or not
related to climate change have the potential to drive extremely complex
changes in community structure and, consequently, food web dynamics.
Developing a greater capability for detecting and understanding these
changes will be critical for future management of ocean and coastal
resources.
30.7.3.9. Socio-ecological Resilience
Many communities depend on marine ecosystems for food and income
yet our understanding of the consequences of environmental degradation
is poor. For example, although there is high confidence that coral reefs
will continue to deteriorate at current rates of climate change and ocean
acidification (Gardner et al., 2003; Bruno and Selig, 2007; De’ath et al.,
2012), there is relatively poor understanding of the implications for the
hundreds of millions of people who depend on these important coastal
ecosystems for food and livelihoods. Improving our understanding of
how to reinforce socio-ecological resilience in communities affected by
the deterioration of key coastal and oceanic ecosystems is central to
developing effective adaptation responses to these growing challenges
(Section 30.6, Tables 30-3, 30-4).
References
Adams, C.M., E. Hernandez, and J.C. Cato, 2004: The economic significance of the
Gulf of Mexico related to population, income, employment, minerals, fisheries
and shipping. Ocean & Coastal Management, 47(11-12), 565-580.
Adjeroud, M., F. Michonneau, P. Edmunds, Y. Chancerelle, T. de Loma, L. Penin, L.
Thibaut, J. Vidal-Dupiol, B. Salvat, and R. Galzin, 2009: Recurrent disturbances,
recovery trajectories, and resilience of coral assemblages on a South Central
Pacific reef. Coral Reefs, 28(3), 775-780.
Al-Rashidi, T.B., H.I. El-Gamily, C.L. Amos, and K.A. Rakha, 2009: Sea surface
temperature trends in Kuwait Bay, Arabian Gulf. Natural Hazards, 50(1), 73-82.
Albright, R. and C. Langdon, 2011: Ocean acidification impacts multiple early life
history processes of the Caribbean coral Porites astreoides. Global Change
Biology, 17(7), 2478-2487.
Albright, R., B. Mason, M. Miller, and C. Langdon, 2010: Ocean acidification compromises
recruitment success of the threatened Caribbean coral Acropora palmata.
Proceedings of the National Academy of Sciences of the United States of
America, 107(47), 20400-20404.
Aleksandrov, S.V., N.N. Zhigalova, and A.S. Zezera, 2009: Long-term dynamics of
zooplankton in the southeastern Baltic Sea. Russian Journal of Marine Biology,
35(4), 296-304.
Alheit, J., 2009: Consequences of regime shifts for marine food webs. International
Journal of Earth Sciences, 98(2), 261-268.
Alheit, J. and A. Bakun, 2010: Population synchronies within and between ocean
basins: apparent teleconnections and implications as to physical–biological
linkage mechanisms. Journal of Marine Systems, 79(3-4), 267-285.
Allen, R. and A. Bergin, 2009: Ocean acidification: an emerging Australian environmental
security challenge. Australian Journal of Maritime and Ocean Affairs, 1(2), 49-
56.
Allison, E.H., A.L. Perry, M.C. Badjeck, W.N. Adger, K. Brown, D. Conway, A.S. Halls,
G.M. Pilling, J.D. Reynolds, and N.L. Andrew, 2009: Vulnerability of national
economies to the impacts of climate change on fisheries. Fish and Fisheries,
10(2), 173-196.
Alonso-Pérez, S., E. Cuevas, and X. Querol, 2011: Objective identification of synoptic
meteorological patterns favouring African dust intrusions into the marine
boundary layer of the subtropical Eastern North Atlantic Region. Meteorology
and Atmospheric Physics, 113(3-4), 109-124.
Amelung, B., S. Nicholls, and D. Viner, 2007: Implications of global climate change
for tourism flows and seasonality. Journal of Travel Research, 45(3), 285-296.
Andersson, A.J. and D. Gledhill, 2013: Ocean acidification and coral reefs: effects on
breakdown, dissolution, and net ecosystem calcification. Annual Review of
Marine Science, 5, 321-348.
Andrews, O.D., N.L. Bindoff, P.R. Halloran, T. Ilyina, and C. Le Quéré, 2013: Detecting
an external influence on recent changes in oceanic oxygen using an optimal
fingerprinting method. Biogeosciences, 10(3), 1799-1813.
Antonov, J.I., D. Seidov, T.P. Boyer, R.A. Locarnini, A.V. Mishonov, H.E. Garcia, O.K.
Baranova, M.M. Zweng, and D.R. Johnson, 2010: World Ocean Atlas 2009,
Volume 2: Salinity. In: NOAA Atlas NESDIS 69 [Levitus, S. (ed.)]. U.S. Government
Printing Office, Washington, DC, USA, 184 pp.
Archer, D., B. Buffett, and V. Brovkin, 2009: Ocean methane hydrates as a slow tipping
point in the global carbon cycle. Proceedings of the National Academy of
Sciences of the United States of America, 106(49), 20596-20601.
Arístegui, J., E.D. Barton, X.A. Álvarez-Salgado, A.M.P. Santos, F.G. Figueiras, S. Kifani,
S. Hernández-León, E. Mason, E. Machú, and H. Demarcq, 2009: Sub-regional
ecosystem variability in the Canary Current upwelling. Progress in Oceanography,
83(1-4), 33-48.
Arrigo, K.R., G. van Dijken, and S. Pabi, 2008: Impact of a shrinking Arctic ice cover
on marine primary production. Geophysical Research Letters, 35(19), L19603,
doi:10.1029/2008GL035028.
Årthun, M., T. Eldevik, L.H. Smedsrud, Ø. Skagseth, and R.B. Ingvaldsen, 2012:
Quantifying the influence of Atlantic heat on Barents Sea ice variability and
retreat. Journal of Climate, 25(13), 4736-4743.
Ashok, K., S.K. Behera, S.A. Rao, H.Y. Weng, and T. Yamagata, 2007: El Niño Modoki
and its possible teleconnection. Journal of Geophysical Research: Oceans,
112(C11), C11007, doi:10.1029/2006JC003798.
Astthorsson, O.S., H. Valdimarsson, A. Gudmundsdottir, and G.J. Óskarsson, 2012:
Climate-related variations in the occurrence and distribution of mackerel
(Scomber scombrus) in Icelandic waters. International Council for the
Exploration of the Sea (ICES) Journal of Marine Science, 69(7), 1289-1297.
Ateweberhan, M. and T.R. McClanahan, 2010: Relationship between historical sea-
surface temperature variability and climate change-induced coral mortality in
the western Indian Ocean. Marine Pollution Bulletin, 60(7), 964-970.
Ateweberhan, M., T.R. McClanahan, N.A.J. Graham, and C.R.C. Sheppard, 2011:
Episodic heterogeneous decline and recovery of coral cover in the Indian
Ocean. Coral Reefs, 30(3), 739-752.
Auad, G., D. Roemmich, and J. Gilson, 2011: The California Current System in relation
to the Northeast Pacific Ocean circulation. Progress in Oceanography, 91, 576-
592.
Ayón, P., S. Purca, and R. Guevara-Carrasco, 2004: Zooplankton volume trends off
Peru between 1964 and 2001. International Council for the Exploration of the
Sea (ICES) Journal of Marine Science, 61(4), 478-484.
Baker, A., P.W. Glynn, and B. Riegl, 2008: Climate change and coral reef bleaching:
an ecological assessment of long-term impacts, recovery trends and future
outlook. Estuarine Coastal and Shelf Science, 80(4), 435-471.
Baker-Austin, C., J.A. Trinanes, N.G.H. Taylor, R. Hartnell, A. Siitonen, and J. Martinez-
Urtaza, 2012: Emerging Vibrio risk at high latitudes in response to ocean
warming. Nature Climate Change, 3(1), 73-77.
Bakun, A., 1990: Global climate change and intensification of coastal ocean
upwelling. Science, 247(4939), 198-201.
Bakun, A., D.B. Field, A. Redondo-Rodriguez, and S.J. Weeks, 2010: Greenhouse
gas, upwelling-favorable winds, and the future of coastal ocean upwelling
ecosystems. Global Change Biology, 16(4), 1213-1228.
Banks, S.A.M.C., S.D. Ling, C.R. Johnson, M.P. Piggott, J.E. Williamson, and L.B.
Beheregaray, 2010: Genetic structure of a recent climate change driven range
extension. Molecular Ecology, 19(10), 2011-2024.
Bañón, R.D., 2009: Variacións na diversidade e abundancia ictiolóxica mariña en
Galicia por efectos do cambio climático. In: Evidencias e Impacto do Cambio
Climático en Galicia. Consellería de Medio Ambiente e Desenvolvemento Sostible,
Xunta de Galicia, Santiago de Compostela, Spain, pp. 355-371 (in Spanish).
Barber, R.T., 1988: Ocean basin ecosystems. In: Concepts of Ecosystem Ecology
[Pomeroy, L.R. and J.J. Alberts (eds.)]. Springer-Verlag, New York, NY, USA, pp.
171-193.
1716
Chapter 30 The Ocean
30
Barber, R.T., J.W. Murray, and J.J. McCarthy, 1994: Biogeochemical interactions in
the equatorial Pacific. AMBIO: A Journal of the Human Environment, 23(1),
62-66.
Barlow, J., M. Kahru, and B. Mitchell, 2008: Cetacean biomass, prey consumption,
and primary production requirements in the California Current ecosystem.
Marine Ecology Progress Series, 371, 285-295.
Barton, A., B. Hales, G.G. Waldbusser, C. Langdon, and R.A. Feely, 2012: The Pacific
oyster, Crassostrea gigas, shows negative correlation to naturally elevated
carbon dioxide levels: implications for near-term ocean acidification effects.
Limnology and Oceanography, 57(3), 698-710.
Barton, E.D., D.B. Field, and C. Roy, 2013: Canary Current upwelling: more or less?
Progress in Oceanography, 116, 167-178.
Batten, S.D. and A.W. Walne, 2011: Variability in northwards extension of warm water
copepods in the NE Pacific. Journal of Plankton Research, 33(11), 1643-1653.
Bauman, A.G., J.A. Burt, D.A. Feary, E. Marquis, and P. Usseglio, 2010: Tropical harmful
algal blooms: an emerging threat to coral reef communities? Marine Pollution
Bulletin, 60(11), 2117-2122.
Beare, D.J., F. Burns, A. Greig, E.G. Jones, K. Peach, M. Kienzle, E. McKenzie, and D.G.
Reid, 2004: Long-term increases in prevalence of North Sea fishes having
southern biogeographic affinities. Marine Ecology Progress Series, 284, 269-
278.
Beare, D., F. Burns, E. Jones, K. Peach, and D. Reid, 2005: Red mullet migration into the
northern North Sea during late winter. Journal of Sea Research, 53(3), 205-212.
Beaudoin, Y. and L. Pendleton (eds.), 2012: Why Value the Oceans? Discussion paper
prepared by the United Nations Environment Programme (UNEP) / Global
Resource Information Database (GRID)-Arendal and Duke University’s
Nicholas Institute for Environmental Policy Solutions in collaboration with
UNEP’s The Economics of Ecosystems and Biodiversity (TEEB) Office and the
UNEP Regional Seas Program, UNEP-TEEB, Geneva, Switzerland, 33 pp.
Beaugrand, G., 2004: The North Sea regime shift: evidence, causes, mechanisms and
consequences. Progress in Oceanography, 60(2), 245-262.
Beaugrand, G., 2009: Decadal changes in climate and ecosystems in the North
Atlantic Ocean and adjacent seas. Deep-Sea Research Part II: Topical Studies
in Oceanography, 56(8-10), 656-673.
Beaugrand, G. and R.R. Kirby, 2010: Spatial changes in the sensitivity of Atlantic
cod to climate-driven effects in the plankton. Climate Research, 41(1), 15-19.
Beaugrand, G. and P.C. Reid, 2003: Long-term changes in phytoplankton, zooplankton
and salmon related to climate. Global Change Biology, 9(6), 801-817.
Beaugrand, G., P.C. Reid, F. Ibanez, J.A. Lindley, and M. Edwards, 2002: Reorganization
of North Atlantic marine copepod biodiversity and climate. Science,
296(5573), 1692-1694.
Beaugrand, G., K.M. Brander, J.A. Lindley, S. Souissi, and P.C. Reid, 2003: Plankton
effect on cod recruitment in the North Sea. Nature, 426(6967), 661-664.
Beaugrand, G., M. Edwards, and L. Legendre, 2010: Marine biodiversity, ecosystem
functioning, and carbon cycles. Proceedings of the National Academy of
Sciences of the United States of America, 107(22), 10120-10124.
Bednaršek, N., G.A. Tarling, D.C.E. Bakker, S. Fielding, E.M. Jones, H.J. Venables, P. Ward,
A. Kuzirian, B. Lézé, R.A. Feely, and E.J. Murphy, 2012: Extensive dissolution
of live pteropods in the Southern Ocean. Nature Geoscience, 5, 881-885.
Belkin, I.M., 2009: Rapid warming of large marine ecosystems. Progress in
Oceanography, 81(1-4), 207-213.
Bell, J.D., J.E. Johnson, and A.J. Hobday (eds.), 2011a: Vulnerability of Tropical Pacific
Fisheries and Aquaculture to Climate Change. Secretariat of the Pacific
Community, Noumea, New Caledonia, 925 pp.
Bell, J.D., C. Reid, M.J. Batty, E.H. Allison, P. Lehodey, L. Rodwell, T.D. Pickering, R. Gillett,
J.E. Johnson, A.J. Hobday, and A. Demmke, 2011b: Implications of climate
change for contributions by fisheries and aquaculture to Pacific Island
economies and communities. In: Vulnerability of Tropical Pacific Fisheries and
Aquaculture to Climate Change [Bell J.D., J.E. Johnson, and A.J. Hobday (eds.)].
Secretariat of the Pacific Community, Noumea, New Caledonia. pp. 733-802.
Bell, J.D., A. Ganachaud, P.C. Gehrke, S.P. Griffiths, A.J. Hobday, O. Hoegh-Guldberg,
J.E. Johnson, R. Le Borgne, P. Lehodey, J.M. Lough, R.J. Matear, T.D. Pickering,
M.S. Pratchett, A. Sen Gupta, I. Senina, and M. Waycott, 2013a: Mixed responses
of tropical Pacific fisheries and aquaculture to climate change. Nature Climate
Change, 3, 591-599.
Bell, J.D., C. Reid, M.J. Batty, P. Lehodey, L. Rodwell, A.J. Hobday, J.E. Johnson, and A.
Demmke, 2013b: Effects of climate change on oceanic fisheries in the tropical
Pacific: implications for economic development and food security. Climatic
Change, 119(1), 199-212.
Ben Rais Lasram, F. and D. Mouillot, 2008: Increasing southern invasion enhances
congruence between endemic and exotic Mediterranean fish fauna. Biological
Invasions, 11(3), 697-711.
Bensoussan, N., J.C. Romano, J.G. Harmelin, and J. Garrabou, 2010: High resolution
characterization of northwest Mediterranean coastal waters thermal regimes:
to better understand responses of benthic communities to climate change.
Estuarine Coastal and Shelf Science, 87(3), 431-441.
Berke, S.K., A.R. Mahon, F.P. Lima, K.M. Halanych, D.S. Wethey, and S.A. Woodin,
2010: Range shifts and species diversity in marine ecosystem engineers:
patterns and predictions for European sedimentary habitats. Global Ecology
and Biogeography, 19(2), 223-232.
Berumen, M.L. and M.S. Pratchett, 2006: Recovery without resilience: persistent
disturbance and long-term shifts in the structure of fish and coral communities
at Tiahura Reef, Moorea. Coral Reefs, 25(4), 647-653.
Bethoux, J., B. Gentili, J. Raunet, and D. Tailliez, 1990: Warming trend in the western
Mediterranean deep water. Nature, 347(6294), 660-662.
Beukema, J.J., R. Dekker, and J.M. Jansen, 2009: Some like it cold: populations of the
tellinid bivalve Macoma balthica (L.) suffer in various ways from a warming
climate. Marine Ecology Progress Series, 384, 135-145.
Bi, H.S., W.T. Peterson, and P.T. Strub, 2011: Transport and coastal zooplankton
communities in the northern California Current system. Geophysical Research
Letters, 38(12), L12607, doi:10.1029/2011GL047927.
Bianchi, C.N., 2007: Biodiversity issues for the forthcoming tropical Mediterranean
Sea. Hydrobiologia, 580(1), 7-21.
Blanchard, J.L., S. Jennings, R. Holmes, J. Harle, G. Merino, J.I. Allen, J. Holt, N.K. Dulvy,
and M. Barange, 2012: Potential consequences of climate change for primary
production and fish production in large marine ecosystems. Philosophical
Transactions of the Royal Society B, 367(1605), 2979-2989.
Block, B.A., I.D. Jonsen, S.J. Jorgensen, A.J. Winship, S.A. Shaffer, S.J. Bograd, E.I. Hazen,
D.G. Foley, G.A. Breed, A.-L. Harrison, J.E. Ganong, A. Swithenbank, M. Castleton,
H. Dewar, B.R. Mate, G.L. Shillinger, K.M. Schaefer, S.R. Benson, M.J. Weise,
R.W. Henry, and D.P. Costa, 2011: Tracking apex marine predator movements
in a dynamic ocean. Nature, 475(7354), 86-90.
Bode, A., R. Anadón, X.A.G. Morán, E. Nogueira, E. Teira, and M. Varela, 2011: Decadal
variability in chlorophyll and primary production off NW Spain. Climate
Research, 48(2), 293-305.
Bograd, S.J., C.G. Castro, E. Di Lorenzo, D.M. Palacios, H. Bailey, W. Gilly, and F.P.
Chavez, 2008: Oxygen declines and the shoaling of the hypoxic boundary in
the California Current. Geophysical Research Letters, 35(12), L12607,
doi:10.1029/2008GL034185.
Bonfils, C. and B.D. Santer, 2011: Investigating the possibility of a human component
in various pacific decadal oscillation indices. Climate Dynamics, 37(7), 1457-
1468.
Bonnet, D., A. Richardson, R. Harris, A. Hirst, G. Beaugrand, M. Edwards, S. Ceballos,
R. Diekman, A. Lopez-Urrutia, L. Valdes, F. Carlotti, J.C. Molinero, H. Weikert,
W. Greve, D. Lucic, A. Albaina, N.D. Yahia, S.F. Umani, A. Miranda, A. dos Santos,
K. Cook, S. Robinson, and M.L.F. de Puelles, 2005: An overview of Calanus
helgolandicus ecology in European waters. Progress in Oceanography, 65(1),
1-53.
Borgerson, S.G., 2008: Arctic meltdown: the economic and security implications of
global warming. Foreign Affairs, 87(2), 63-77.
Bostock, J., B. McAndrew, R. Richards, K. Jauncey, T. Telfer, K. Lorenzen, D. Little, L.
Ross, N. Handisyde, I. Gatward, and R. Corner, 2010: Aquaculture: global status
and trends. Philosophical Transactions of the Royal Society B, 365, 2897-2912.
Boyer, T.P., 2005: Linear trends in salinity for the World Ocean, 1955-1998. Geophysical
Research Letters, 32(1), L01604, doi:10.1029/2004GL021791.
Boyle, A., 2012: Law of the Sea perspectives on climate change. The International
Journal of Marine and Coastal Law, 27(4), 831-838.
Bozkurt, D. and O.L. Sen, 2011: Precipitation in the Anatolian Peninsula: sensitivity
to increased SSTs in the surrounding seas. Climate Dynamics, 36(3-4), 711-
726.
Brevik, E.C., and J.A. Homburg, 2004: A 5000 year record of carbon sequestration
from a coastal lagoon and wetland complex, Southern California, USA. Catena,
57(3), 221-232.
Brown, B.E., R.P. Dunne, N. Phongsuwan, and P.J. Somerfield, 2011: Increased sea
level promotes coral cover on shallow reef flats in the Andaman Sea, eastern
Indian Ocean. Coral Reefs, 30(4), 867-878.
Brown, C.W. and J.A. Yoder, 1994: Coccolithophorid blooms in the global ocean.
Journal of Geophysical Research: Oceans, 99(C4), 7467-7482.
1717
The Ocean Chapter 30
30
Brown, J.H., J.F. Gillooly, A.P. Allen, V.M. Savage, and G.B. West, 2004: Toward a
metabolic theory of ecology. Ecology, 85(7), 1771-1789.
Brundrit, G.B., 1995: Trends in southern African sea level: statistical analysis and
interpretation. South African Journal of Marine Science, 16(1), 9-17.
Bruno, J.F. and E.R. Selig, 2007: Regional decline of coral cover in the Indo-Pacific:
timing, extent, and subregional comparisons. PLoS ONE, 2(8), e711,
doi:10.1371/journal.pone.0000711.
Buesseler, K.O., C.H. Lamborg, P.W. Boyd, P.J. Lam, T.W. Trull, R.R. Bidigare, J.K.B. Bishop,
K.L. Casciotti, F. Dehairs, M. Elskens, M. Honda, D.M. Karl, D.A. Siegel, M.W. Silver,
D.K. Steinberg, J. Valdes, B. Van Mooy, and S. Wilson, 2007: Revisiting carbon
flux through the ocean’s twilight zone. Science, 316(5824), 567-570.
Buongiorno Nardelli, B., S. Colella, R. Santoleri, M. Guarracino, and A. Kholod, 2010:
A re-analysis of Black Sea surface temperature. Journal of Marine Systems,
79(1-2), 50-64.
Burke, L.M. and J. Maidens, 2004: Reefs at Risk in the Caribbean. World Resources
Institute (WRI), Washington, DC, USA, 80 pp.
Burke, L.M., L. Selig, and M. Spalding, 2002: Reefs at Risk in Southeast Asia. World
Resources Institute (WRI), Washington, DC, USA, 76 pp.
Burke, L.M., K. Reytar, M. Spalding, and A. Perry, 2011: Reefs at Risk Revisited. World
Resources Institute (WRI), Washington, DC, USA, 130 pp.
Burrows, M.T., D.S. Schoeman, L.B. Buckley, P. Moore, E.S. Poloczanska, K.M. Brander,
C. Brown, J.F. Bruno, C.M. Duarte, B.S. Halpern, J. Holding, C.V. Kappel, W.
Kiessling, M.I. O’Connor, J.M. Pandolfi, C. Parmesan, F.B. Schwing, W.J.
Sydeman, and A.J. Richardson, 2011: The pace of shifting climate in marine
and terrestrial ecosystems. Science, 334(6056), 652-655.
Busby, J.W., 2009: The climate-security connection: what it means for the poor. In:
Climate Change and Global Poverty: A Billion Lives in the Balance [Brainard,
L., A. Jones, and N. Purvis (eds.)]. Brookings Institute Press, Washington, DC,
USA, pp. 155-180.
Byrne, R.H., S. Mecking, R.A. Feely, and X. Liu, 2010: Direct observations of basin-
wide acidification of the North Pacific Ocean. Geophysical Research Letters,
37(2), L02601, doi:10.1029/2009GL040999.
Cai, R. and H. Tan, 2010: Influence of interdecadal climate variation over East Asia
on offshore ecological system of China. Journal of Oceanography in Taiwan
Strait, 29(2), 173-183 (in Chinese).
Cai, R., J. Chen, and R. Huang, 2006: The response of marine environment in the
offshore area of China and its adjacent ocean to recent global climate change.
Chinese Journal of Atmospheric Sciences, 30(5), 1019-1033 (in Chinese).
Cai, R., Q. Zhang, and Q. Qi, 2009: Spatial and temporal oscillation and long-term
variation in sea surface temperature field of the South China Sea. Journal of
Oceanography in Taiwan Strait, 29(2), 173-183 (in Chinese).
Cai, R., J. Chen, and H. Tan, 2011: Variations of the sea surface temperature in the
offshore area of China and their relationship with the East Asian monsoon
under the global warming. Climatic and Environmental Research, 16(1), 94-
104 (in Chinese).
Cai, W., 2006: Antarctic ozone depletion causes an intensification of the Southern
Ocean super-gyre circulation. Geophysical Research Letters, 33(3), L03712,
doi:10.1029/2005GL024911.
Cai, W. and T. Cowan, 2007: Trends in Southern Hemisphere circulation in IPCC AR4
models over 1950-99: ozone depletion versus greenhouse forcing. Journal of
Climate, 20(4), 681-693.
Cai, W., M. Lengaigne, S. Borlace, M. Collins, T. Cowan, M.J. McPhaden, A. Timmermann,
S. Power, J. Brown, C. Menkes, A. Ngari, E.M. Vincent, and M.J. Widlansky, 2012:
More extreme swings of the South Pacific convergence zone due to greenhouse
warming. Nature, 488(7411), 365-369.
Caldeira, K. and M.E. Wickett, 2003: Oceanography: anthropogenic carbon and ocean
pH. Nature, 425(6956), 365, doi:10.1038/425365a .
Callaghan, J. and S.B. Power, 2011: Variability and decline in the number of severe
tropical cyclones making land-fall over eastern Australia since the late
nineteenth century. Climate Dynamics, 37(3), 647-662.
Calvo, E., R. Simó, R. Coma, M. Ribes, J. Pascual, A. Sabatés, J.M. Gili, and C. Pelejero,
2011: Effects of climate change on Mediterranean marine ecosystems: the
case of the Catalan Sea. Climate Research, 50(1), 1-29.
Cantin, N.E., A.L. Cohen, K.B. Karnauskas, A.M. Tarrant, and D.C. McCorkle, 2010:
Ocean warming slows coral growth in the central Red Sea. Science,
329(5989), 322-325.
Carrigan, A.D. and M.L. Puotinen, 2011: Assessing the potential for tropical cyclone
induced sea surface cooling to reduce thermal stress on the world’s coral reefs.
Geophysical Research Letters, 38(23), L23604, doi:10.1029/2011GL049722.
Casini, M., G. Kornilovs, M. Cardinale, C. Möllmann, W. Grygiel, P. Jonsson, T. Raid, J.
Flinkman, and V. Feldman, 2011: Spatial and temporal density dependence
regulates the condition of the central Baltic Sea clupeids: compelling evidence
using an extensive international acoustic survey. Population Ecology, 53(4),
511-523.
Cerrano, C., G. Bavestrello, C. Bianchi, R. Cattaneo-Vietti, S. Bava, C. Morganti, C.
Morri, P. Picco, G. Sara, and S. Schiaparelli, 2000: A catastrophic mass-mortality
episode of gorgonians and other organisms in the Ligurian Sea (North-western
Mediterranean), summer 1999. Ecology Letters, 3(4), 284-293.
Chan, F., J.A. Barth, J. Lubchenco, A. Kirincich, H. Weeks, W.T. Peterson, and B.A.
Menge, 2008: Emergence of anoxia in the California Current large marine
ecosystem. Science, 319(5865), 920, doi:10.1126/science.1149016.
Chan, W.H., Y.L. Mak, J.J. Wu, L. Jin, W.H. Sit, J.C. Lam, Y. Sadovy de Mitcheson, L.L.
Chan, P.K. Lam, and M.B. Murphy, 2011: Spatial distribution of ciguateric fish
in the Republic of Kiribati. Chemosphere, 84(1), 117-123.
Chang, Y., M.-A. Lee, K.-T. Lee, and K.-T. Shao, 2013: Adaptation of fisheries and
ariculture management to extreme oceanic environmental changes and
climate variability in Taiwan. Marine Policy, 38, 476-482.
Chassot, E., F. Mélin, O. Le Pape, and D. Gascuel, 2007: Bottom-up control regulates
fisheries production at the scale of eco-regions in European seas. Marine
Ecology Progress Series, 343, 45-55.
Chassot, E., S. Bonhommeau, N.K. Dulvy, F. Mélin, R. Watson, D. Gascuel, and O. Le
Pape, 2010: Global marine primary production constrains fisheries catches.
Ecology Letters, 13(4), 495-505.
Chavez, F.P. and M. Messie, 2009: A comparison of Eastern Boundary Upwelling
Ecosystems. Progress in Oceanography, 83(1-4), 80-96.
Chavez, F.P., J. Ryan, S.E. Lluch-Cota, and C.M. Ñiquen, 2003: From anchovies to
sardines and back: multidecadal change in the Pacific Ocean. Science,
299(5604), 217-221.
Chavez, F.P., M. Messie, and J.T. Pennington, 2011: Marine primary production in
relation to climate variability and change. Annual Review of Marine Science,
3, 227-260.
Checkley, D.M. and J.A. Barth, 2009: Patterns and processes in the California Current
System. Progress in Oceanography, 83(1-4), 49-64.
Chelton, D.B., P.A. Bernal, and J.A. McGowan, 1982: Large-scale interannual physical
and biological interaction in the California Current. Journal of Marine Research,
40(4), 1095-1125.
Chen, T.-R., K.-F. Yu, Q. Shi, S. Li, G.J. Price, R. Wang, M.-X. Zhao, T.-G. Chen, and J.-X.
Zhao, 2009: Twenty-five years of change in scleractinian coral communities
of Daya Bay (northern South China Sea) and its response to the 2008 AD
extreme cold climate event. Chinese Science Bulletin, 54(12), 2107-2117.
Cheng, X. and Y. Qi, 2007: Trends of sea level variations in the South China Sea from
merged altimetry data. Global and Planetary Change, 57(3-4), 371-382.
Cheung, W.W.L., V.W.Y. Lam, J.L. Sarmiento, K. Kearney, R. Watson, D. Zeller, and D.
Pauly, 2010: Large-scale redistribution of maximum fisheries catch potential
in the global ocean under climate change. Global Change Biology, 16(1), 24-
35.
Cheung, W.W.L., J. Dunne, J.L. Sarmiento, and D. Pauly, 2011: Integrating
ecophysiology and plankton dynamics into projected maximum fisheries catch
potential under climate change in the Northeast Atlantic. International Council
for the Exploration of the Sea (ICES) Journal of Marine Science, 68(6), 1008-
1018.
Cheung, W.W.L., J.K. Pinnegar, G. Merino, M.C. Jones, and M. Barange, 2012: Review
of climate change impacts on marine fisheries in the UK and Ireland. Aquatic
Conservation: Marine and Freshwater Ecosystems, 22, 368-388.
Chhak, K.C., E. Di Lorenzo, N. Schneider, and P.F. Cummins, 2009: Forcing of low-
frequency ocean variability in the Northeast Pacific. Journal of Climate, 22(5),
1255-1276.
Chiba, S., K. Tadokoro, H. Sugisaki, and T. Saino, 2006: Effects of decadal climate
change on zooplankton over the last 50 years in the western subarctic North
Pacific. Global Change Biology, 12(5), 907-920.
Chiba, S., M.N. Aita, K. Tadokoro, T. Saino, H. Sugisaki, and K. Nakata, 2008: From
climate regime shifts to lower-trophic level phenology: synthesis of recent
progress in retrospective studies of the western North Pacific. Progress in
Oceanography, 77(2-3), 112-126.
China-SNAP, 2011: The Second National Chinese Assessment Report on Climate
Change: General Abstract. Jointly issued by the Chinese Ministry of Science
and Technology, the China Meteorological Administration, and the Chinese
Academy of Sciences, Science Press, Beijing, China, 44 pp.
1718
Chapter 30 The Ocean
30
Christian, J.R. and R. Murtugudde, 2003: Tropical Atlantic variability in a coupled
physical–biogeochemical ocean model. Deep-Sea Research Part II: Topical
Studies in Oceanography, 50(22-26), 2947-2969.
Church, J.A., N.J. White, R. Coleman, K. Lambeck, and J.X. Mitrovica, 2004: Estimates
of the regional distribution of sea level rise over the 1950-2000 period. Journal
of Climate, 17(13), 2609-2625.
Church, J.A., N.J. White, and J.R. Hunter, 2006: Sea-level rise at tropical Pacific and
Indian Ocean islands. Global and Planetary Change, 53(3), 155-168.
CIESM, 2008: Climate warming and related changes in Mediterranean marine biota.
In: CIESM Workshop Monographs, No. 35 [Briand, F. (ed.)]. Commission
Internationale pour l’Exploration Scientifique de la Méditerranée (CIESM) [The
Mediterranean Science Commission], Monaco, 152 pp.
Cinner, J.E., T.R. McClanahan, T.M. Daw, N.A. Graham, J. Maina, S.K. Wilson, and T.P.
Hughes, 2009: Linking social and ecological systems to sustain coral reef
fisheries. Current Biology, 19(3), 206-212.
Clark, C.O., J.E. Cole, and P.J. Webster, 2000: Indian Ocean SST and Indian summer
rainfall: predictive relationships and their decadal variability. Journal of
Climate, 13(14), 2503-2519.
Clifton, J., 2009: Science, funding and participation: key issues for marine protected
area networks and the Coral Triangle Initiative. Environmental Conservation,
36(2), 91-96.
Cocco, V., F. Joos, M. Steinacher, T.L. Frölicher, L. Bopp, J. Dunne, M. Gehlen, C. Heinze,
J. Orr, and A. Oschlies, 2013: Oxygen and indicators of stress for marine life in
multi-model global warming projections. Biogeosciences, 10(3), 1849-1868.
Cochrane, K.L., N.L. Andrew, and A.M. Parma, 2011: Primary fisheries management:
a minimum requirement for provision of sustainable human benefits in small-
scale fisheries. Fish and Fisheries, 12(3), 275-288.
Cockcroft, A.C., D. van Zyl, and L. Hutchings, 2008: Large-scale changes in the spatial
distribution of South African West Coast rock lobsters: an overview. African
Journal of Marine Science, 30(1), 149-159.
Collie, J.S., A.D. Wood, and H.P. Jeffries, 2008: Long-term shifts in the species
composition of a coastal fish community. Canadian Journal of Fisheries and
Aquatic Sciences, 65(7), 1352-1365.
Collins, M., S.-I. An, W. Cai, A. Ganachaud, E. Guilyardi, F.-F. Jin, M. Jochum, M.
Lengaigne, S. Power, and A. Timmermann, 2010: The impact of global warming
on the tropical Pacific Ocean and El Niño. Nature Geoscience, 3(6), 391-397.
Colwell, R.R., 1996: Global climate and infectious disease: the cholera paradigm.
Science, 274(5295), 2025-2031.
Coma, R., M. Ribes, E. Serrano, E. Jiménez, J. Salat, and J. Pascual, 2009: Global
warming-enhanced stratification and mass mortality events in the
Mediterranean. Proceedings of the National Academy of Sciences of the
United States of America, 106(15), 6176-6181.
Conley, D.J., S. Bjorck, E. Bonsdorff, J. Carstensen, G. Destouni, B.G. Gustafsson, S.
Hietanen, M. Kortekaas, H. Kuosa, H.E. Meier, B. Muller-Karulis, K. Nordberg,
A. Norkko, G. Nurnberg, H. Pitkanen, N.N. Rabalais, R. Rosenberg, O.P. Savchuk,
C.P. Slomp, M. Voss, F. Wulff, and L. Zillen, 2009: Hypoxia-related processes in
the Baltic Sea. Environmental Science & Technology, 43(10), 3412-3420.
Conversi, A., S. Fonda Umani, T. Peluso, J.C. Molinero, A. Santojanni, and M. Edwards,
2010: The Mediterranean Sea regime shift at the end of the 1980s, and
intriguing parallelisms with other European basins. PLoS ONE, 5(5), e10633,
doi:10.1371/journal.pone.0010633.
Cooley, S.R. and S.C. Doney, 2009: Anticipating ocean acidification’s economic
consequences for commercial fisheries. Environmental Research Letters, 4(2),
024007, doi:10.1088/1748-9326/4/2/024007.
Costello, A., M. Abbas, A. Allen, S. Ball, S. Bell, R. Bellamy, S. Friel, N. Groce, A. Johnson,
M. Kett, M. Lee, C. Levy, M. Maslin, D. McCoy, B. McGuire, H. Montgomery, D.
Napier, C. Pagel, J. Patel, J.A. de Oliveira, N. Redclift, H. Rees, D. Rogger, J. Scott,
J. Stephenson, J. Twigg, J. Wolff, and C. Patterson, 2009: Managing the health
effects of climate change. Lancet, 373(9676), 1693-1733.
Crain, C.M., K. Kroeker, and B.S. Halpern, 2008: Interactive and cumulative effects of
multiple human stressors in marine systems. Ecology Letters, 11(12), 1304-1315.
Crawford, R.J.M., A.J. Tree, P.A. Whittington, J. Visagie, L. Upfold, K.J. Roxburg, A.P.
Martin, and B.M. Dyer, 2008: Recent distributional changes of seabirds in
South Africa: is climate having an impact? African Journal of Marine Science,
30(1), 189-193.
Crisci, C., N. Bensoussan, J.C. Romano, and J. Garrabou, 2011: Temperature anomalies
and mortality events in marine communities: insights on factors behind
differential mortality impacts in the NW Mediterranean. PLoS ONE, 6(9),
e23814, doi:10.1371/journal.pone.0023814.
Cruz, A.M. and E. Krausmann, 2008: Damage to offshore oil and gas facilities following
hurricanes Katrina and Rita: an overview. Journal of Loss Prevention in the
Process Industries, 21(6), 620-626.
Cubillos, J.C., S.W. Wright, G. Nash, M.F. de Salas, B. Griffiths, B. Tilbrook, A. Poisson,
and G.M. Hallegraeff, 2007: Calcification morphotypes of the coccolithophorid
Emiliania huxleyi in the Southern Ocean: changes in 2001 to 2006 compared
to historical data. Marine Ecology Progress Series, 348, 47-54.
Cury, P. and L. Shannon, 2004: Regime shifts in upwelling ecosystems: observed
changes and possible mechanisms in the northern and southern Benguela.
Progress in Oceanography, 60(2-4), 223-243.
Cushing, D.H., 1990: Plankton production and year-class strength in fish populations
an update of the match/mismatch hypothesis. In: Advances in Marine Biology,
Vol. 26 [Blaxter, J.H.S. and A.J. Southward (eds.)]. Academic Press, San Diego,
CA, USA and London, UK, pp. 249-293.
Dalpadado, P., R.B. Ingvaldsen, L.C. Stige, B. Bogstad, T. Knutsen, G. Ottersen, and
B. Ellertsen, 2012: Climate effects on Barents Sea ecosystem dynamics.
International Council for the Exploration of the Sea (ICES) Journal of Marine
Science, 69(7), 1303-1316.
Danovaro, R., S. Fonda Umani, and A. Pusceddu, 2009: Climate change and the potential
spreading of marine mucilage and microbial pathogens in the Mediterranean
Sea. PLoS ONE, 4(9), e7006, doi:10.1371/journal.pone.0007006.
Danovaro, R., C. Corinaldesi, G. D’Onghia, B. Galil, C. Gambi, A.J. Gooday, N.
Lampadariou, G.M. Luna, C. Morigi, K. Olu, P. Polymenakou, E. Ramirez-Llodra,
A. Sabbatini, F. Sardà, M. Sibuet, and A. Tselepides, 2010: Deep-sea biodiversity
in the Mediterranean Sea: the known, the unknown, and the unknowable.
PLoS ONE, 5(8), e11832, doi:10.1371/journal.pone.0011832.
De’ath, G., K.E. Fabricius, H. Sweatman, and M. Puotinen, 2012: The 27-year decline
of coral cover on the Great Barrier Reef and its causes. Proceedings of the
National Academy of Sciences of the United States of America, 109(44),
17995-17999.
Della Patrona, L., B. Beliaeff, and T. Pickering, 2011: Mitigation of sea level rise effects
by addition of sediment to shrimp ponds. Aquaculture Environment Interactions,
2, 27-38.
Demarcq, H., 2009: Trends in primary production, sea surface temperature and wind
in upwelling systems (1998-2007). Progress in Oceanography, 83(1), 376-385.
Deser, C., A.S. Phillips, and J.W. Hurrell, 2004: Pacific interdecadal climate variability:
linkages between the tropics and the North Pacific during boreal winter since
1900. Journal of Climate, 17(16), 3109-3124.
Di Lorenzo, E., N. Schneider, K.M. Cobb, P.J.S. Franks, K. Chhak, A.J. Miller, J.C.
McWilliams, S.J. Bograd, H. Arango, E. Curchitser, T.M. Powell, and P. Rivière,
2008: North Pacific Gyre Oscillation links ocean climate and ecosystem change.
Geophysical Research Letters, 35(8), L08607, doi:10.1029/2007GL032838.
Diaz-Almela, E., N. Marbà, and C.M. Duarte, 2007: Consequences of Mediterranean
warming events in seagrass (Posidonia oceanica) flowering records. Global
Change Biology, 13(1), 224-235.
Dixson, D.L., G.P. Jones, P.L. Munday, S. Planes, M.S. Pratchett, M. Srinivasan, C. Syms,
and S.R. Thorrold, 2008: Coral reef fish smell leaves to find island homes.
Proceedings of the Royal Society B, 275(1653), 2831-2839.
Dixson, D.L., P.L. Munday, and G.P. Jones, 2010: Ocean acidification disrupts the
innate ability of fish to detect predator olfactory cues. Ecology Letters, 13(1),
68-75.
Doney, S.C., V.J. Fabry, R.A. Feely, and J.A. Kleypas, 2009: Ocean acidification: the
other CO
2
problem. Annual Review of Marine Science, 1, 169-192.
Donner, S.D., C.J. Kucharik, and J.A. Foley, 2004: Impact of changing land use
practices on nitrate export by the Mississippi River. Global Biogeochemical
Cycles, 18(1), GB1028, doi:10.1029/2003GB002093.
Donner, S.D., W.J. Skirving, C.M. Little, M. Oppenheimer, and O. Hoegh-Guldberg,
2005: Global assessment of coral bleaching and required rates of adaptation
under climate change. Global Change Biology, 11(12), 2251-2265.
Donner, S.D., T.R. Knutson, and M. Oppenheimer, 2007: Model-based assessment
of the role of human-induced climate change in the 2005 Caribbean coral
bleaching event. Proceedings of the National Academy of Sciences of the
United States of America, 104(13), 5483-5488.
Donner, S.D., T. Kirata, and C. Vieux, 2010: Recovery from the 2004 coral bleaching
event in the Gilbert Islands, Kiribati. Atoll Research Bulletin, 587, 1-27.
Dove, S., D.I. Kline, O. Pantos, F.E. Angly, G.W. Tyson, and O. Hoegh-Guldberg, 2013:
Future reef decalcification under a business-as-usual CO
2
emission scenario.
Proceedings of the National Academy of Sciences of the United States of
America, 110(38), 15342-15347.
1719
The Ocean Chapter 30
30
Drinkwater, K.F., 2006: The regime shift of the 1920s and 1930s in the North Atlantic.
Progress in Oceanography, 68(2-4), 134-151.
Duarte, C.M., J. Middelburg, and N. Caraco, 2005: Major role of marine vegetation
on the oceanic carbon cycle. Biogeosciences, 2(1), 1-8.
Dulvy, N.K., S.I. Rogers, S. Jennings, V. Stelzenmüller, S.R. Dye, and H.R. Skjoldal, 2008:
Climate change and deepening of the North Sea fish assemblage: a biotic
indicator of warming seas. Journal of Applied Ecology, 45(4), 1029-1039.
Durack, P.J. and S.E. Wijffels, 2010: Fifty-year trends in global ocean salinities and their
relationship to broad-scale warming. Journal of Climate, 23(16), 4342-4362.
Durack, P.J., S.E. Wijffels, and R.J. Matear, 2012: Ocean salinities reveal strong global
water cycle intensification during 1950 to 2000. Science, 336(6080), 455-458.
Eakin, C.M., J.A. Morgan, S.F. Heron, T.B. Smith, G. Liu, L. Alvarez-Filip, B. Baca, E.
Bartels, C. Bastidas, C. Bouchon, M. Brandt, A.W. Bruckner, L. Bunkley-Williams,
A. Cameron, B.D. Causey, M. Chiappone, T.R.L. Christensen, M.J.C. Crabbe, O.
Day, E. de la Guardia, G. Díaz-Pulido, D. DiResta, D.L. Gil-Agudelo, D.S. Gilliam,
R.N. Ginsburg, S. Gore, H.M. Guzmán, J.C. Hendee, E.A. Hernández-Delgado,
E. Husain, C.F.G. Jeffrey, R.J. Jones, E. Jordán-Dahlgren, L.S. Kaufman, D.I. Kline,
P.A. Kramer, J.C. Lang, D. Lirman, J. Mallela, C. Manfrino, J.-P. Maréchal, K.
Marks, J. Mihaly, W.J. Miller, E.M. Mueller, E.M. Muller, C.A. Orozco Toro, H.A.
Oxenford, D. Ponce-Taylor, N. Quinn, K.B. Ritchie, S. Rodríguez, A.R. Ramírez,
S. Romano, J.F. Samhouri, J.A. Sánchez, G.P. Schmahl, B.V. Shank, W.J. Skirving,
S.C.C. Steiner, E. Villamizar, S.M. Walsh, C. Walter, E. Weil, E.H. Williams, K.W.
Roberson, and Y. Yusuf, 2010: Caribbean corals in crisis: record thermal stress,
bleaching, and mortality in 2005. PLoS ONE, 5(11), e13969, doi:10.1371/
journal.pone.0013969.
Edgar, G.J., S.A. Banks, M. Brandt, R.H. Bustamante, A. Chiriboga, S.A. Earle, L.E.
Garske, P.W. Glynn, J.S. Grove, S. Henderson, C.P. Hickman, K.A. Miller, F. Rivera,
and G.M. Wellington, 2010: El Niño, grazers and fisheries interact to greatly
elevate extinction risk for Galapagos marine species. Global Change Biology,
16(10), 2876-2890.
Edwards, M., 2004: Continuous plankton records: Plankton Atlas of the North
Atlantic Ocean 1958-1999: Preface. Marine Ecology Progress Series, (Suppl.
CPR Atlas), 1-2, www.int-res.com/journals/meps/cpr-plankton-atlas-2004/.
Edwards, M. and A.J. Richardson, 2004: Impact of climate change on marine pelagic
phenology and trophic mismatch. Nature, 430(7002), 881-884.
Edwards, M., G. Beaugrand, G.C. Hays, J.A. Koslow, and A.J. Richardson, 2010: Multi-
decadal oceanic ecological datasets and their application in marine policy and
management. Trends in Ecology & Evolution, 25(10), 602-610.
Ellingsen, I.H., P. Dalpadado, D. Slagstad, and H. Loeng, 2008: Impact of climatic
change on the biological production in the Barents Sea. Climatic Change,
87(1), 155-175.
Emanuel, K., 2005: Increasing destructiveness of tropical cyclones over the past 30
years. Nature, 436(7051), 686-688.
Eriksen, E., R. Ingvaldsen, J.E. Stiansen, and G.O. Johansen, 2012: Thermal habitat
for 0-group fish in the Barents Sea; how climate variability impacts their
density, length, and geographic distribution. International Council for the
Exploration of the Sea (ICES) Journal of Marine Science, 69(5), 870-879.
Fabricius, K.E., C. Langdon, S. Uthicke, C. Humphrey, S. Noonan, G. De’ath, R. Okazaki,
N. Muehllehner, M.S. Glas, and J.M. Lough, 2011: Losers and winners in coral
reefs acclimatized to elevated carbon dioxide concentrations. Nature Climate
Change, 1(3), 165-169.
Falkowski, P.G., T. Algeo, L. Codispoti, C. Deutsch, S. Emerson, B. Hales, R.B. Huey,
W.J. Jenkins, L.R. Kump, L.A. Levin, T.W. Lyons, N.B. Nelson, O.S. Schofield, R.
Summons, L.D. Talley, E. Thomas, F. Whitney, and C.B. Pilcher, 2011: Ocean
deoxygenation: past, present, and future. Eos, 92(46), 409-410.
FAO, 2012. The State of the World Fisheries and Aquaculture 2012. Food and
Agriculture Organization of the United Nations (FAO), Rome, Italy, 209 pp.
Feely, R.A., C.L. Sabine, K. Lee, W. Berelson, J. Kleypas, V.J. Fabry, and F.J. Millero,
2004: Impact of anthropogenic CO
2
on the CaCO
3
system in the oceans.
Science, 305(5682), 362-366.
Feely, R.A., C.L. Sabine, J.M. Hernandez-Ayon, D. Ianson, and B. Hales, 2008: Evidence
for upwelling of corrosive “acidified” water onto the continental shelf. Science,
320(5882), 1490-1492.
Feely, R.A., S.C. Doney, and S.R. Cooley, 2009: Ocean acidification: present conditions
and future changes in a high-CO
2
world. Oceanography, 22(4), 36-47.
Feely, R.A., C.L. Sabine, R.H. Byrne, F.J. Millero, A.G. Dickson, R. Wanninkhof, A.
Murata, L.A. Miller, and D. Greeley, 2012: Decadal changes in the aragonite
and calcite saturation state of the Pacific Ocean. Global Biogeochemical
Cycles, 26(3), GB3001, doi:10.1029/2011GB004157.
Firth, L.B., A.M. Knights, and S.S. Bell, 2011: Air temperature and winter mortality:
implications for the persistence of the invasive mussel, Perna viridis in the
intertidal zone of the south-eastern United States. Journal of Experimental
Marine Biology and Ecology, 400(1-2), 250-256.
Fonselius, S. and J. Valderrama, 2003: One hundred years of hydrographic
measurements in the Baltic Sea. Journal of Sea Research, 49(4), 229-241.
Form, A.U. and U. Riebesell, 2012: Acclimation to ocean acidification during long-
term CO
2
exposure in the cold-water coral Lophelia pertusa. Global Change
Biology, 18(3), 843-853.
Friedrich, T., A. Timmermann, A. Abe-Ouchi, N.R. Bates, M.O. Chikamoto, M.J. Church,
J.E. Dore, D.K. Gledhill, M. Gonzalez-Davila, M. Heinemann, T. Ilyina, J.H.
Jungclaus, E. McLeod, A. Mouchet, and J.M. Santana-Casiano, 2012: Detecting
regional anthropogenic trends in ocean acidification against natural variability.
Nature Climate Change, 2(3), 167-171.
Frieler, K., M. Meinshausen, A. Golly, M. Mengel, K. Lebek, S.D. Donner, and O. Hoegh-
Guldberg, 2012: Limiting global warming to 2°C is unlikely to save most coral
reefs. Nature Climate Change, 3, 165-170.
Frontier Economics, Ltd., 2013: Economics of Climate Resilience, Natural Environment
Theme: Sea Fish, CA0401. Report prepared for the Department for Environment
Food and Rural Affairs (DEFRA) and the Devolved Administrations (DAs) by
Frontier Economics, Ltd. supported by Irbaris LLP and Ecofys UK, Ltd., DEFRA,
London, UK, 101 pp.
Fu, B.-J., C.-H. Su, Y.-P. Wei, I.R. Willett, Y.-H. Lü, and G.-H. Liu, 2011: Double counting
in ecosystem services valuation: causes and countermeasures. Ecological
Research, 26(1), 1-14.
Galil, B.S., 2008: Alien species in the Mediterranean Sea – which, when, where, why?
Hydrobiologia, 606(1), 105-116.
Galil, B.S., 2011: The alien crustaceans in the Mediterranean Sea: an historical review.
In: In the Wrong Place –Alien Marine Crustaceans: Distribution, Biology and
Impacts [Galil, B.S., P.F. Clark, and J.T. Carlton (eds.)]. Springer, Dordrecht,
Netherlands, pp. 377-401.
Ganachaud, A., A. Sen Gupta, J. Orr, S. Wijffels, K. Ridgway, M. Hemer, C. Maes, C.
Steinberg, A. Tribollet, B. Qiu, and K. J., 2011: Observed and expected changes
to the tropical Pacific Ocean. In: Vulnerability of Tropical Pacific Fisheries and
Aquaculture to Climate Change [Bell, J.D., J.E. Johnson, and A.J. Hobday
(eds.)]. Secretariat of the Pacific Community, Noumea, New Caledonia. pp.
101-187.
Gardner, T.A., I.M. Côté, J.A. Gill, A. Grant, and A.R. Watkinson, 2003: Long-term
region-wide declines in Caribbean corals. Science, 301(5635), 958-960.
Garrabou, J., R. Coma, N. Bensoussan, M. Bally, P. Chevaldonné, M. Cigliano, D. Diaz,
J.G. Harmelin, M.C. Gambi, D.K. Kersting, J.B. Ledoux, C. Lejeusne, C. Linares,
C. Marschal, T. Pérez, M. Ribes, J.C. Romano, E. Serrano, N. Teixido, O. Torrents,
M. Zabala, F. Zuberer, and C. Cerrano, 2009: Mass mortality in Northwestern
Mediterranean rocky benthic communities: effects of the 2003 heat wave.
Global Change Biology, 15(5), 1090-1103.
Garreaud, R.D. and M. Falvey, 2009: The coastal winds off western subtropical South
America in future climate scenarios. International Journal of Climatology,
29(4), 543-554.
Gattuso, J.-P., J. Bijma, M. Gehlen, U. Riebesell, and C. Turley, 2011: Ocean acidification:
knowns, unknowns, and perspectives. In: Ocean Acidification [Gattuso, J.-P.
and L. Hansson (eds.)]. Oxford University Press, Oxford, UK, pp. 291-311.
Genner, M.J., D.W. Sims, A.J. Southward, G.C. Budd, P. Masterson, M. McHugh, P.
Rendle, E.J. Southall, V.J. Wearmouth, and S.J. Hawkins, 2010: Body size-
dependent responses of a marine fish assemblage to climate change and
fishing over a century-long scale. Global Change Biology, 16(2), 517-527.
Geresi, E., R. Chapman, T. McGee, and B. Woolsey, 2009: Monitoring seafloor
instability caused by the presence of gas hydrate using ocean acoustical and
geophysical techniques in the northern Gulf of Mexico. In: Natural Gas
Hydrates Energy Resource Potential and Associated Geologic Hazards
[Collett, T., A. Johnson, C. Knapp, and R. Boswell (eds.)]. AAPG Memoir 89, Co-
Published by The American Association of Petroleum Geologists (AAPG),
Energy Minerals Division, with the U.S. Department of Energy, National Energy
Technology Laboratory (NETL), and the AAPG Foundation, AAPG, Tulsa, OK,
USA, pp. 287-296.
Gillett, N.P. and P.A. Stott, 2009: Attribution of anthropogenic influence on seasonal
sea level pressure. Geophysical Research Letters, 36(23), L23709, doi:10.1029/
2009GL041269.
Gillett, N.P., F.W. Zwiers, A.J. Weaver, and P.A. Stott, 2003: Detection of human
influence on sea-level pressure. Nature, 422(6929), 292-294.
1720
Chapter 30 The Ocean
30
Gilman, E., J. Ellison, and R. Coleman, 2007: Assessment of mangrove response to
projected relative sea-level rise and recent historical reconstruction of shoreline
position. Environmental Monitoring and Assessment, 124(1-3), 105-130.
Gleckler, P.J., B. Santer, C. Domingues, D. Pierce, T. Barnett, J. Church, K. Taylor, K.
AchutaRao, T. Boyer, and M. Ishii, 2012: Human-induced global ocean warming
on multidecadal timescales. Nature Climate Change, 2, 524-529.
Gledhill, D.K., R. Wanninkhof, F.J. Millero, and M. Eakin, 2008: Ocean acidification of
the Greater Caribbean Region 1996-2006. Journal of Geophysical Research,
113(C10), C10031, doi:10.1029/2007JC004629.
Glynn, P.W., 1984: Widespread coral mortality and the 1982-83 El Niño warming
event. Environmental Conservation, 11(2), 133-146.
Glynn, P.W., 2001: Eastern Pacific coral reef ecosystems. In: Coastal Marine Ecosystems
of Latin America [Seeliger, U. and B. Kjerfve (eds.)]. Ecological Studies, Vol. 144,
Springer-Verlag, Berlin, Germany, pp. 281-305.
Glynn, P.W., 2011: In tandem reef coral and cryptic metazoan declines and extinctions.
Bulletin of Marine Science, 87(4), 767-794.
Gómez-Gesteira, M., M. De Castro, I. Álvarez, M.N. Lorenzo, J.L.G. Gesteira, and A.J.C.
Crespo, 2008: Spatio-temporal upwelling trends along the Canary Upwelling
System (1967-2006). Annals of the New York Academy of Sciences, 1146(1),
320-337.
González-Taboada, F. and R. Anadón, 2012: Patterns of change in sea surface
temperature in the North Atlantic during the last three decades: beyond mean
trends. Climatic Change, 115(2), 419-431.
Goreau, T., T. McClanahan, R. Hayes, and A. Strong, 2000: Conservation of coral reefs
after the 1998 global bleaching event. Conservation Biology, 14(1), 5-15.
Graham, N.A., S.K. Wilson, S. Jennings, N.V. Polunin, J. Robinson, J.P. Bijoux, and T.M.
Daw, 2007: Lag effects in the impacts of mass coral bleaching on coral reef
fish, fisheries, and ecosystems. Conservation Biology, 21(5), 1291-1300.
Graham, N.A., T.R. McClanahan, M.A. MacNeil, S.K. Wilson, N.V. Polunin, S. Jennings,
P. Chabanet, S. Clark, M.D. Spalding, Y. Letourneur, L. Bigot, R. Galzin, M.C.
Öhman, K.C. Garpe, A.J. Edwards, and C.R. Sheppard, 2008: Climate warming,
marine protected areas and the ocean-scale integrity of coral reef ecosystems.
PLoS ONE, 3(8), e3039, doi:10.1371/journal.pone.0003039.
Grantham, B.A., F. Chan, K.J. Nielsen, D.S. Fox, J.A. Barth, A. Huyer, J. Lubchenco,
and B.A. Menge, 2004: Upwelling-driven nearshore hypoxia signals ecosystem
and oceanographic changes in the northeast Pacific. Nature, 429(6993), 749-
754.
Green, D., L. Alexander, K. Mclnnes, J. Church, N. Nicholls, and N. White, 2010: An
assessment of climate change impacts and adaptation for the Torres Strait
Islands, Australia. Climatic Change, 102(3-4), 405-433.
Gregg, W.W., N.W. Casey, and C.R. McClain, 2005: Recent trends in global ocean
chlorophyll. Geophysical Research Letters, 32(3), L03606, doi:10.1029/
2004GL021808.
Griffith, G.P., E.A. Fulton, R. Gorton, and A.J. Richardson, 2012: Predicting interactions
among fishing, ocean warming, and ocean acidification in a marine system
with whole-ecosystem models. Conservation Biology, 26(6), 1145-1152.
Gruber, N., 2011: Warming up, turning sour, losing breath: ocean biogeochemistry
under global change. Philosophical Transactions of the Royal Society A,
369(1943), 1980-1996.
Gruber, N., C. Hauri, Z. Lachkar, D. Loher, T.L. Frölicher, and G.K. Plattner, 2012: Rapid
progression of ocean acidification in the California Current System. Science,
337(6091), 220-223.
Guilyardi, E., H. Bellenger, M. Collins, S. Ferrett, W. Cai, and A. Wittenberg, 2012: A
first look at ENSO in CMIP5. CLIVAR Exchanges No.58, 17(1), 29-32.
Guinotte, J.M., J. Orr, S. Cairns, A. Freiwald, L. Morgan, and R. George, 2006: Will
human-induced changes in seawater chemistry alter the distribution of
deep-sea scleractinian corals? Frontiers in Ecology and the Environment, 4(3),
141-146.
Gutiérrez, D., I. Bouloubassi, A. Sifeddine, S. Purca, K. Goubanova, M. Graco, D. Field,
L. Méjanelle, F. Velazco, A. Lorre, R. Salvatteci, D. Quispe, G. Vargas, B. Dewitte,
and L. Ortlieb, 2011: Coastal cooling and increased productivity in the main
upwelling zone off Peru since the mid-twentieth century. Geophysical
Research Letters, 38, L07603, doi:10.1029/2010GL046324.
Hall-Spencer, J.M., R. Rodolfo-Metalpa, S. Martin, E. Ransome, M. Fine, S.M. Turner,
S.J. Rowley, D. Tedesco, and M.-C. Buia, 2008: Volcanic carbon dioxide vents
show ecosystem effects of ocean acidification. Nature, 454(7200), 96-99.
Halpern, B.S., K.L. McLeod, A.A. Rosenberg, and L.B. Crowder, 2008: Managing for
cumulative impacts in ecosystem-based management through ocean zoning.
Ocean & Coastal Management, 51(3), 203-211.
Halpern, B.S., C.M. Ebert, C.V. Kappel, E.M.P. Madin, F. Micheli, M. Perry, K.A. Selkoe,
and S. Walbridge, 2009: Global priority areas for incorporating land-sea
connections in marine conservation. Conservation Letters, 2(4), 189-196.
Halpern, B.S., S.E. Lester, and K.L. McLeod, 2010: Placing marine protected areas
onto the ecosystem-based management seascape. Proceedings of the National
Academy of Sciences of the United States of America, 107(43), 18312-18317.
Hänninen, J. and I. Vuorinen, 2011: Time-varying parameter analysis of the Baltic
Sea freshwater runoffs. Environmental Modeling & Assessment, 16(1), 53-60.
Hare, S.R. and N.J. Mantua, 2000: Empirical evidence for North Pacific regime shifts
in 1977 and 1989. Progress in Oceanography, 47(2-4), 103-145.
Harley, C.D., 2011: Climate change, keystone predation, and biodiversity loss. Science,
334(6059), 1124-1127.
Harvell, C.D., K. Kim, J.M. Burkholder, R.R. Colwell, P.R. Epstein, D.J. Grimes, E.E.
Hofmann, E.K. Lipp, A.D. Osterhaus, R.M. Overstreet, J.W. Porter, G.W. Smith, and
G.R. Vasta, 1999: Emerging marine diseases-climate links and anthropogenic
factors. Science, 285(5433), 1505-1510.
Harvell, C.D., C.E. Mitchell, J.R. Ward, S. Altizer, A.P. Dobson, R.S. Ostfeld, and M.D.
Samuel, 2002: Climate warming and disease risks for terrestrial and marine
biota. Science, 296(5576), 2158-2162.
Harvell, D., R. Aronson, N. Baron, J. Connell, A. Dobson, S. Ellner, L. Gerber, K. Kim, A.
Kuris, H. McCallum, K. Lafferty, B. McKay, J. Porter, M. Pascual, G. Smith, K.
Sutherland, and J. Ward, 2004: The rising tide of ocean diseases: unsolved
problems and research priorities. Frontiers in Ecology and the Environment,
2(7), 375-382.
Hashioka, T., T.T. Sakamoto, and Y. Yamanaka, 2009: Potential impact of global
warming on North Pacific spring blooms projected by an eddy-permitting 3-D
ocean ecosystem model. Geophysical Research Letters, 36, L20604, doi:10.1029/
2009GL038912.
Hátún, H., M. Payne, G. Beaugrand, P. Reid, A. Sandø, H. Drange, B. Hansen, J. Jacobsen,
and D. Bloch, 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(3-4), 149-162.
Hazen, E.L., S. Jorgensen, R.R. Rykaczewski, S.J. Bograd, D.G. Foley, I.D. Jonsen, S.A.
Shaffer, J.P. Dunne, D.P. Costa, and L.B. Crowder, 2012: Predicted habitat shifts
of Pacific top predators in a changing climate. Nature Climate Change, 3, 234-
238.
Head, E.J.H. and P. Pepin, 2010: Monitoring changes in phytoplankton abundance
and composition in the Northwest Atlantic: a comparison of results obtained
by continuous plankton recorder sampling and colour satellite imagery.
Journal of Plankton Research, 32(12), 1649-1660.
Healy, T. and K. Harada, 1991: Definition and physical characteristics of the world’s
enclosed coastal seas. Marine Pollution Bulletin, 23, 639-644.
Hegerl, G.C., F.W. Zwiers, P. Braconnot, N.P. Gillett, Y. Luo, J.A. Marengo Orsini, N.
Nicholls, J.T. Penner, and P. Stott, 2007: Understanding and attributing climate
change. In: Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [S. Solomon, D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor, and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, UK and New York, NY, USA, pp. 643-745.
Hegerl, G.C., O. Hoegh-Guldberg, G. Casassa, M.P. Hoerling, R.S. Kovats, C. Parmesan,
D.W. Pierce, and P.A. Stott, 2010: Good practice guidance paper on detection
and attribution related to anthropogenic climate change. In: Meeting Report of
the Intergovernmental Panel on Climate Change Expert Meeting on Detection
and Attribution of Anthropogenic Climate Change [Stocker, T.F., C.B. Field, D.
Qin, V. Barros, G.-K. Plattner, M. Tignor, P.M. Midgley, and K.L. Ebi (eds.)]. IPCC
Working Group I Technical Support Unit, University of Bern, Bern, Switzerland,
8 pp.
Heil, C.A., P.M. Glibert, M.A. Al-Sarawi, M. Faraj, M. Behbehani, and M. Husain, 2001:
First record of a fish-killing Gymnodinium sp. bloom in Kuwait Bay, Arabian Sea:
chronology and potential causes. Marine Ecology Progress Series, 214, 15-23.
Heiss, G.A., 1996: Annual band width variation in Porites sp. from Aqaba, Gulf of
Aqaba, Red Sea. Bulletin of Marine Science, 59(2), 393-403.
Heitmann, N. and S. Khalilian, 2010: Accounting for CO
2
Emissions from International
Shipping: Burden Sharing Under Different UNFCCC Allocation Options and
Regime Scenarios. Kiel Working Paper No. 1655, Kiel Institute for the World
Economy, Kiel, Germany, 36 pp.
Helly, J.J. and L.A. Levin, 2004: Global distribution of naturally occurring marine
hypoxia on continental margins. Deep-Sea Research Part I: Oceanographic
Research Papers, 51(9), 1159-1168.
1721
The Ocean Chapter 30
30
Helm, K.P., N.L. Bindoff, and J.A. Church, 2011: Observed decreases in oxygen content
of the global ocean. Geophysical Research Letters, 38(23), L23602, doi:10.1029/
2011GL049513.
Henriksen, P., 2009: Long-term changes in phytoplankton in the Kattegat, the Belt
Sea, the Sound and the western Baltic Sea. Journal of Sea Research, 61(1-2),
114-123.
Henson, S.A., J.L. Sarmiento, J.P. Dunne, L. Bopp, I.D. Lima, S.C. Doney, J. John, and C.
Beaulieu, 2010: Detection of anthropogenic climate change in satellite records
of ocean chlorophyll and productivity. Biogeosciences, 7, 621-640.
Hermansen, Ø. and K. Heen, 2012: Norwegian salmonid farming and global
warming: socioeconomic impacts. Aquaculture Economics & Management,
16(3), 220-221.
Hermant, M., J. Lobry, S. Bonhommeau, J.C. Poulard, and O. Le Pape, 2010: Impact
of warming on abundance and occurrence of flatfish populations in the Bay
of Biscay (France). Journal of Sea Research, 64(1-2), 45-53.
Heymans, J.J., L.J. Shannon, and A. Jarre, 2004: Changes in the northern Benguela
ecosystem over three decades: 1970s, 1980s, and 1990s. Ecological Modelling,
172(2-4), 175-195.
Hickey, B.M., 1979: The California Current System – hypotheses and facts. Progress
in Oceanography, 8(4), 191-279.
Hiddink, J.G. and R. ter Hofstede, 2008: Climate induced increases in species richness
of marine fishes. Global Change Biology, 14(3), 453-460.
Hilborn, R., 2007: Defining success in fisheries and conflicts in objectives. Marine
Policy, 31(2), 153-158.
Hill, K.L., S.R. Rintoul, R. Coleman, and K.R. Ridgway, 2008: Wind forced low frequency
variability of the East Australia Current. Geophysical Research Letters, 35(8),
L08602, doi:10.1029/2007GL032912.
Hilmi, N., D. Allemand, S. Dupont, A. Safa, G. Haraldsson, P.A.L.D. Nunes, C. Moore, C.
Hattam, S. Reynaud, J.M. Hall-Spencer, M. Fine, C. Turley, R. Jeffree, J. Orr, P.L.
Munday, and S.R. Cooley, 2013: Towards improved socio-economic assessments
of ocean acidification’s impacts. Marine Biology, 160(8), 1773-1787.
Hitchcock, G.L., P. Lane, S. Smith, J.G. Luo, and P.B. Ortner, 2002: Zooplankton spatial
distributions in coastal waters of the northern Arabian Sea, August, 1995.
Deep-Sea Research Part II: Topical Studies in Oceanography, 49(12), 2403-
2423.
Hjort, J., 1914: Fluctuations in the great fisheries of Northern Europe. Rapports et
Procès-Verbaux des Reunions du Conseil Permanent International pour
l’Exploration de la Mer (CIEM) [International Council for the Exploration of
the Sea (ICES)], 20, 1-228.
Hoegh-Guldberg, O., 1999: Climate change, coral bleaching and the future of the
world’s coral reefs. Marine and Freshwater Research, 50(8), 839-866.
Hoegh-Guldberg, O., 2011: Coral reef ecosystems and anthropogenic climate
change. Regional Environmental Change, 11(1 Suppl.), S215-S227.
Hoegh-Guldberg, O., 2012: The adaptation of coral reefs to climate change: is the
Red Queen being outpaced? Scientia Marina, 76(2), 403-408.
Hoegh-Guldberg, O. and J.F. Bruno, 2010: The impact of climate change on the
world’s marine ecosystems. Science, 328(5985), 1523-1528.
Hoegh-Guldberg, O. and B. Salvat, 1995: Periodic mass-bleaching and elevated sea
temperatures – bleaching of outer reef slope communities in Moorea, French
Polynesia. Marine Ecology Progress Series, 121(1-3), 181-190.
Hoegh-Guldberg, O., P.J. Mumby, A.J. Hooten, R.S. Steneck, P. Greenfield, E. Gomez,
C.D. Harvell, P.F. Sale, A.J. Edwards, K. Caldeira, N. Knowlton, C.M. Eakin, R.
Iglesias-Prieto, N. Muthiga, R.H. Bradbury, A. Dubi, and M.E. Hatziolos, 2007:
Coral reefs under rapid climate change and ocean acidification. Science,
318(5857), 1737-1742.
Hoegh-Guldberg, O., L. Hughes, S. McIntyre, D.B. Lindenmayer, C. Parmesan, H.P.
Possingham, and C.D. Thomas, 2008: Assisted colonization and rapid climate
change. Science, 321(5887), 345-346.
Hoegh-Guldberg, O., H. Hoegh-Guldberg, J.E.N. Veron, A. Green, E.D. Gomez, J.M.
Lough, M. King, A. Ambariyanto, L. Hansen, J.E. Cinner, G. Dews, G. Russ, H.Z.
Shuttenberg, E.L. Penaflor, C.M. Eakin, T.R.L. Christiensen, M. Abbey, F. Areki,
R.A. Kosaka, A. Tewfik, and J. Oliver, 2009: The Coral Triangle and Climate
Change: Ecosystems, People and Societies at Risk. World Wide Fund For Nature
(WWF) Australia, Brisbane, QLD, Australia, 276 pp.
Hoegh-Guldberg, O., S. Andréfouët, K.E. Fabricius, G. Diaz-Pulido, J.M. Lough, P.A.
Marshall, and M.S. Pratchett, 2011a: Vulnerability of coral reefs in the tropical
Pacific to climate change. In: Vulnerability of Tropical Pacific Fisheries and
Aquaculture to Climate Change [Bell, J.D., J.E. Johnson, and A.J. Hobday (eds.)].
Secretariat of the Pacific Community, Noumea, New Caledonia. pp. 251-296.
Hoegh-Guldberg, O., G. Hegerl, T. Root, F. Zwiers, P. Stott, D. Pierce, and M. Allen,
2011b: Correspondence: difficult but not impossible. Nature Climate Change,
1(2), 72, doi:10.1038/nclimate1107.
Hoegh-Guldberg, O., T. Aqorau, R. Arnason, T. Chansiri, N. Del Rio, H. Demone, S.
Earle, M.H. Feeley, D. Gutierrez, R. Hilborn, N. Ishil, C. Lischewski, J. Lubchenco,
K. Anh Nguyen, D. Obura, H.E.R. Payet, T. Neroni Slade, J. Tanzer, J.H. Williams,
D.J. Wright, and J. Xu, 2013: Indispensable Ocean: Aligning Ocean Health and
Human Well-Being Guidance from the Blue Ribbon Panel to the Global
Partnership for Oceans. Working Paper Report No. 82651, The World Bank,
Washington, DC USA, 40 pp.
Hofmann, A.F., E.T. Peltzer, and P.G. Brewer, 2013: Kinetic bottlenecks to respiratory
exchange rates for deep-sea – Part I: Oxygen. Biogeosciences, 10, 5049-5060.
Hönisch, B., A. Ridgwell, D.N. Schmidt, E. Thomas, S.J. Gibbs, A. Sluijs, R.E. Zeebe, L.
Kump, R.C. Martindale, S.E. Greene, W. Kiessling, J. Ries, J.C. Zachos, D.L. Royer,
S. Barker, T.M. Marchitto Jr., R. Moyer, C. Pelejero, P. Zivera, G.L. Foster, and B.
Williams, 2012: The geological record of ocean acidification. Science,
335(6072), 1058-1063.
Hosoda, S., T. Suga, N. Shikama, and K. Mizuno, 2009: Global surface layer salinity
change detected by Argo and its implication for hydrological cycle intensification.
Journal of Oceanography, 65, 579-586.
Howard, J.A., A. Jarre, A. Clark, and C. Moloney, 2007: Application of the sequential
t-test algorithm for analysing regime shifts to the southern Benguela ecosystem.
African Journal of Marine Science, 29(3), 437-451.
Howell, E.A., C.C.C. Wabnitz, J.P. Dunne, and J.J. Polovina, 2013: Climate-induced
primary productivity change and fishing impacts on the Central North Pacific
ecosystem and Hawaii-based pelagic longline fishery. Climatic Change,
119(1), 79-93.
Hughes, T.P., 1994: Catastrophes, phase shifts, and large-scale degradation of a
Caribbean coral reef. Science, 265(5178), 1547-1551.
Hughes, T.P., A.H. Baird, D.R. Bellwood, M. Card, S.R. Connolly, C. Folke, R. Grosberg,
O. Hoegh-Guldberg, J.B. Jackson, J. Kleypas, J.M. Lough, P. Marshall, M. Nystrom,
S.R. Palumbi, J.M. Pandolfi, B. Rosen, and J. Roughgarden, 2003: Climate change,
human impacts, and the resilience of coral reefs. Science, 301(5635), 929-933.
Hughes, T.P., D.R. Bellwood, A.H. Baird, J. Brodie, J.F. Bruno, and J.M. Pandolfi, 2011:
Shifting base-lines, declining coral cover, and the erosion of reef resilience:
comment on Sweatman et al. (2011). Coral Reefs, 30(3), 653-660.
Hunt, G.L., P. Stabeno, G. Walters, E. Sinclair, R.D. Brodeur, J.M. Napp, and N.A. Bond,
2002: Climate change and control of the southeastern Bering Sea pelagic
ecosystem. Deep-Sea Research Part II: Topical Studies in Oceanography,
49(26), 5821-5853.
Hunt, G.L., B.M. Allen, R.P. Angliss, T. Baker, N. Bond, G. Buck, G.V. Byrd, K.O. Coyle, A.
Devol, D.M. Eggers, L. Eisner, R.A. Feely, S. Fitzgerald, L.W. Fritz, E.V. Gritsay,
C. Ladd, W. Lewis, J.T. Mathis, C.W. Mordy, F. Mueter, J. Napp, E. Sherr, D. Shull,
P. Stabeno, M.A. Stephanenko, S. Strom, and T.E. Whitledge, 2010: Status and
trends of the Bering Sea region, 2003-2008. In: Marine Ecosystems of the
North Pacific Ocean, 2003-2008 [McKinnell, S.M. and M.J. Dagg (eds.)]. PICES
Special Publication No. 4, North Pacific Marine Science Organization (PICES),
Sydney, BC, Canada, pp. 196-267.
Huse, G. and I. Ellingsen, 2008: Capelin migrations and climate change a modelling
analysis. Climatic Change, 87(1), 177-197.
Huse, G., J.C. Holst, K. Utne, L. Nøttestad, W. Melle, A. Slotte, G. Ottersen, T. Fenchel,
and F. Uiblein, 2012: Effects of interactions between fish populations on
ecosystem dynamics in the Norwegian Sea – results of the INFERNO project.
Marine Biology Research, 8(5-6), 415-419.
Hutchings, L., C.D. van der Lingen, L.J. Shannon, R.J.M. Crawford, H.M.S. Verheye,
C.H. Bartholomae, A.K. van der Plas, D. Louw, A. Kreiner, M. Ostrowski, Q. Fidel,
R.G. Barlow, T. Lamont, J. Coetzee, F. Shillington, J. Veitch, J.C. Currie, and P.M.S.
Monteiro, 2009: The Benguela Current: an ecosystem of four components.
Progress in Oceanography, 83(1-4), 15-32.
ICES, 2012: Book 3: The Barents Sea and the Norwegian Sea. In: ICES Advice 2012,
Books 1-10. Report of the ICES Advisory Committee 2012, International Council
for the Exploration of the Sea (ICES), Copenhagen, Denmark, 79 pp.
Iles, A.C., T.C. Gouhier, B.A. Menge, J.S. Stewart, A.J. Haupt, and M.C. Lynch, 2012:
Climate-driven trends and ecological implications of event-scale upwelling in
the California Current System. Global Change Biology, 18(2), 783-796.
IPCC, 2005: IPCC Special Report on Carbon Dioxide Capture and Storage. Prepared
by Working Group III of the Intergovernmental Panel on Climate Change (IPCC)
[Metz, B., O. Davidson, H.C. de Coninck, M. Loos, and L.A. Meyer (eds.)].
Cambridge University Press, Cambridge, UK and New York, NY, USA, 442 pp.
1722
Chapter 30 The Ocean
30
IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate
Change Adaptation. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change [Field, C.B., V. Barros, T.F. Stocker,
D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K.
Allen, M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press,
Cambridge, UK and New York, NY, USA, 582 pp.
Irvine, J.R. and M.A. Fukuwaka, 2011: Pacific salmon abundance trends and climate
change. International Council for the Exploration of the Sea (ICES) Journal of
Marine Science, 68(6), 1122-1130.
Ito, S., H. Sugisaki, A. Tsuda, O. Yamamura, and K. Okuda, 2004: Contributions of the
VENFISH program: meso-zooplankton, Pacific saury (Cololabis saira) and
walleye pollock (Theragra chalcogramma) in the northwestern Pacific. Fisheries
Oceanography, 13(Suppl. s1), 1-9.
Ito, S., K.A. Rose, A.J. Miller, K. Drinkwater, K. Brander, J.E. Overland, S. Sundby, E.
Curchitser, J.W. Hurrell, and Y. Yamanaka, 2010: Ocean ecosystem responses
to future global change scenarios: a way forward. In: Marine Ecosystems and
Global Change [Barange, M., J.G. Field, R.H. Harris, E. Hofmann, R.I. Perry, and
F. Werner (eds.)]. Oxford University Press, Oxford, UK, pp. 287-322.
Ittekkot, V., S. Jilan, E. Miles, E. Desa, B.N. Desai, J.T. Everett, J.J. Magnuson, A. Tsyban,
and S. Zuta, 1996: Oceans. In: Climate Change 1995: Impacts, Adaptations,
and Mitigation of Climate Change: Scientific-Technical Analyses. Contribution
of Working Group II to the Second Assessment Report of the Intergovernmental
Panel on Climate Change [Watson, R.T., M. Zinyowera, and R. Moss (eds.)].
Cambridge University Press, Cambridge, UK and New York, NY, USA, pp. 267-288.
Iversen, T., 1934: Some observations on cod in northern waters. In: Report on Norwegian
Fishery and Marine Investigations, Vol.4 No. 8. Norwegian Directorate of
Fisheries and John Griegs Boktrykkeri, Bergen, Norway, pp. 1-35, brage.
bibsys.no/xmlui/bitstream/handle/11250/114626/1/sh_vol04_08_1934.pdf.
Jenkins, W.J., 2008: The biogeochemical consequences of changing ventilation in the
Japan/East Sea. Marine Chemistry, 108(3-4), 137-147.
Jin, M.B., C. Deal, J. Wang, and C.P. McRoy, 2009: Response of lower trophic level
production to long-term climate change in the southeastern Bering Sea.
Journal of Geophysical Research, 114, C04010, doi:10.1029/2008JC005105.
Johannesen, E., R.B. Ingvaldsen, B. Bogstad, P. Dalpadado, E. Eriksen, H. Gjøsæter, T.
Knutsen, M. Skern-Mauritzen, and J.E. Stiansen, 2012: Changes in Barents Sea
ecosystem state, 1970-2009: climate fluctuations, human impact, and trophic
interactions. International Council for the Exploration of the Sea (ICES) Journal
of Marine Science, 69(5), 880-889.
Jordà, G., D. Gomis, E. Alvarez-Fanjul, and S. Somot, 2012: Atmospheric contribution
to Mediterranean and nearby Atlantic sea level variability under different
climate change scenarios. Global and Planetary Change, 80-81, 198-214.
Jung, S., 2008: Spatial variability in long-term changes of climate and oceanographic
conditions in Korea. Journal of Environmental Biology, 29(4), 519-529.
Jung, S., I.-C. Pang, J.-H. Lee, I. Choi, and H.K. Cha, 2014: Latitudinal shifts in the
distribution of exploited fishes in Korean waters during the last 30 years: a
consequence of climate change. Reviews in Fisheries Biology and Fisheries,
24, 443-462, doi:10.1007/s11160-013-9310-1.
Jury, M., T. McClanahan, and J. Maina, 2010: West Indian Ocean variability and East
African fish catch. Marine Environmental Research, 70(2), 162-170.
Justić, D., N.N. Rabalais, and R.E. Turner, 1996: Effects of climate change on hypoxia
in coastal waters: a doubled CO
2
scenario for the northern Gulf of Mexico.
Limnology and Oceanography, 41(5), 992-1003.
Justić, D., V.J. Bierman, D. Scavia, and R.D. Hetland, 2007: Forecasting Gulf’s hypoxia:
the next 50 years? Estuaries and Coasts, 30(5), 791-801.
Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, L. Gandin, M. Iredell, S.
Saha, G. White, J. Woollen, Y. Zhu, A. Leetmaa, R. Reynolds, M. Chelliah, W.
Ebisuzaki, W. Higgins, J. Janowiak, K.C. Mo, C. Ropelewski, J. Wang, R. Jenne,
and D. Joseph, 1996: The NCEP/NCAR 40-year reanalysis project. Bulletin of
the American Meteorological Society, 77(3), 437-471.
Kane, J., 2007: Zooplankton abundance trends on Georges Bank, 1977-2004.
International Council for the Exploration of the Sea (ICES) Journal of Marine
Science, 64(5), 909-919.
Karstensen, J., L. Stramma, and M. Visbeck, 2008: Oxygen minimum zones in the
eastern tropical Atlantic and Pacific oceans. Progress in Oceanography, 77(4),
331-350.
Kaye, S., 2012: Climate change and maritime security. In: Climate Change and the
Oceans: Gauging the Legal and Policy Currents in the Asia Pacific and Beyond
[Warner, R. and C. Schofield (eds.)]. Edward Elgar Publishing, Cheltenham, UK
and Northampton, MA, USA, pp. 153-166.
Keeling, R.E., A. Körtzinger, and N. Gruber, 2010: Ocean deoxygenation in a warming
world. Annual Review of Marine Science, 2, 199-229.
Keister, J.E., E. Di Lorenzo, C.A. Morgan, V. Combes, and W.T. Peterson, 2011:
Zooplankton species composition is linked to ocean transport in the Northern
California Current. Global Change Biology, 17(7), 2498-2511.
Kelly, R.P., M.M. Foley, W.S. Fisher, R.A. Feely, B.S. Halpern, G.G. Waldbusser, and R.R.
Caldwell, 2011: Mitigating local causes of ocean acidification with existing
laws. Science, 332(6033), 1036-1037.
Kenchington, R. and R. Warner, 2012: Uncertain seas ahead: legal and policy
approaches to conserving marine biodiversity in the face of a changing
climate. In: Climate Change and the Oceans: Gauging the Legal and Policy
Currents in the Asia Pacific and Beyond [Warner, R. and C. Schofield (eds.)].
Edward Elgar Publishing, Cheltenham, UK and Northampton, MA, USA, pp.
153-166.
Kerr, R.A., 2000: A North Atlantic climate pacemaker for the centuries. Science,
288(5473), 1984-1985.
Kessler, J.D., D.L. Valentine, M.C. Redmond, M. Du, E.W. Chan, S.D. Mendes, E.W.
Quiroz, C.J. Villanueva, S.S. Shusta, L.M. Werra, S.A. Yvon-Lewis, and T.C. Weber,
2011: A persistent oxygen anomaly reveals the fate of spilled methane in the
deep Gulf of Mexico. Science, 331(6015), 312-315.
Key, R.M., A. Kozyr, C.L. Sabine, K. Lee, R. Wanninkhof, J.L. Bullister, R.A. Feely, F.J.
Millero, C.W. Mordy, and T.H. Peng, 2004: A global ocean carbon climatology:
results from Global Data Analysis Project (GLODAP). Global Biogeochemical
Cycles, 18(4), GB4031, doi:10.1029/2004GB002247.
Kideys, A.E., 2002: Fall and rise of the Black Sea ecosystem. Science, 297(5586),
1482-1484.
King, J.R., V.N. Agostini, C.J. Harvey, G.A. McFarlane, M.G.G. Foreman, J.E. Overland,
E. Di Lorenzo, N.A. Bond, and K.Y. Aydin, 2011: Climate forcing and the
California Current ecosystem. International Council for the Exploration of the
Sea (ICES) Journal of Marine Science, 68(6), 1199-1216.
Kirby, R.R. and G. Beaugrand, 2009: Trophic amplification of climate warming.
Proceedings of the Royal Society B, 276(1676), 4095-4103.
Kirkman, S.P., D. Yemane, J. Kathena, S.K. Mafwila, S.E. Nsiangango, T. Samaai,
B. Axelsen, and L. Singh, 2013: Identifying and characterizing demersal
fish biodiversity hotspots in the Benguela Current Large Marine
Ecosystem: relevance in the light of global changes. International Council for
the Exploration of the Sea (ICES) Journal of Marine Science, 70(5), 943-
954.
Kleypas, J.A., R.W. Buddemeier, D. Archer, J.-P. Gattuso, C. Langdon, and B.N. Opdyke,
1999: Geochemical consequences of increased atmospheric carbon dioxide
on coral reefs. Science, 284(5411), 118-120.
Knoll, A.H. and W.W. Fischer, 2011: Skeletons and ocean chemistry: the long view.
In: Ocean Acidification [Gattuso, J.-P. and L. Hansson (eds.)]. Oxford University
Press, Oxford, UK, pp. 67-82.
Knutson, T.R., J.L. McBride, J. Chan, K. Emanuel, G. Holland, C. Landsea, I. Held, J.P.
Kossin, A. Srivastava, and M. Sugi, 2010: Tropical cyclones and climate change.
Nature Geoscience, 3(3), 157-163.
Koslow, J.A., R. Goericke, A. Lara-Lopez, and W. Watson, 2011: Impact of declining
intermediate-water oxygen on deepwater fishes in the California Current.
Marine Ecology Progress Series, 436, 207-218.
Krishnan, P., S.D. Roy, G. George, R.C. Srivastava, A. Anand, S. Murugesan, M.
Kaliyamoorthy, N. Vikas, and R. Soundararajan, 2011: Elevated sea surface
temperature during May 2010 induces mass bleaching of corals in the
Andaman. Current Science, 100(1), 111-117.
Kroeker, K.J., R.L. Kordas, R. Crim, I.E. Hendriks, L. Ramajo, G.S. Singh, C.M. Duarte,
and J.-P. Gattuso, 2013: Impacts of ocean acidification on marine organisms:
quantifying sensitivies and interaction with warming. Global Change Biology,
19(6), 1884-1896.
Kurien, J. and R. Willmann, 2009: Special considerations for small-scale fisheries
management in developing countries. In: A Fishery Manager’s Guidebook
[Cochrane, K. and S.M. Garcia (eds.)]. Food and Agricultural Organization of
the United Nations (FAO), Rome, Italy and Wiley-Blackwell, Chichester, UK, pp.
404-424.
Kuzmina, S.I., L. Bengtsson, O.M. Johannessen, H. Drange, L.P. Bobylev, and M.W. Miles,
2005: The North Atlantic Oscillation and greenhouse-gas forcing. Geophysical
Research Letters, 32(4), L04703, doi:10.1029/2004GL021064.
Laffoley, D. and G.D. Grimsditch (eds.), 2009: The Management of Natural Coastal
Carbon Sinks. International Union for Conservation of Nature and Natural
Resources (IUCN), Gland, Switzerland, 53 pp.
1723
The Ocean Chapter 30
30
Last, P.R., W.T. White, D.C. Gledhill, A.J. Hobday, R. Brown, G.J. Edgar, and G. Pecl,
2011: Long-term shifts in abundance and distribution of a temperate fish
fauna: a response to climate change and fishing practices. Global Ecology and
Biogeography, 20(1), 58-72.
Laws, E.A., P.G. Falkowski, W.O. Smith, H. Ducklow, and J.J. McCarthy, 2000: Temperature
effects on export production in the open ocean. Global Biogeochemical Cycles,
14(4), 1231-1246.
Le Borgne, R., R.T. Barber, T. Delcroix, H.Y. Inoue, D.J. Mackey, and M. Rodier, 2002:
Pacific warm pool and divergence: temporal and zonal variations on the equator
and their effects on the biological pump. Deep-Sea Research Part II: Topical
Studies in Oceanography, 49(13-14), 2471-2512.
Le Borgne, W.R., V. Allain, S.P. Griffiths, R.J. Matear, A.D. McKinnon, A.J. Richardson,
and J.W. Young, 2011: Vulnerability of open oceans food webs in the tropical
Pacific to climate change. In: Vulnerability of Tropical Pacific Fisheries and
Aquaculture to Climate Change [Bell, J.D., J.E. Johnson, and A.J. Hobday (eds.)].
Secretariat of the Pacific Community, Noumea, New Caledonia, pp. 189-250.
Leduc, G., C.T. Herbert, T. Blanz, P. Martinez, and R. Schneider, 2010: Contrasting
evolution of sea surface temperature in the Benguela upwelling system under
natural and anthropogenic climate forcings. Geophysical Research Letters,
37(20), L20705, doi:10.1029/2010GL044353.
Lee, S.-K., W. Park, E. van Sebille, M.O. Baringer, C. Wang, D.B. Enfield, S.G. Yeager,
and B.P. Kirtman, 2011: What caused the significant increase in Atlantic Ocean
heat content since the mid-20
t
h
century? Geophysical Research Letters, 38(17),
L17607, doi:10.1029/2011GL048856.
Lee, S.-S., J.-Y. Lee, B. Wang, K.-J. Ha, K.-Y. Heo, F.-F. Jin, D.M. Straus, and J. Shukla,
2012: Interdecadal changes in the storm track activity over the North Pacific
and North Atlantic. Climate Dynamics, 39(1-2), 313-327.
Lee, T. and M.J. McPhaden, 2010: Increasing intensity of El Niño in the central-
equatorial Pacific. Geophysical Research Letters, 37(14), L14603, doi:10.1029/
2010GL044007.
Lee, Z., S. Shang, C. Hu, M. Lewis, R. Arnone, Y. Li, and B. Lubac, 2010: Time series of
bio-optical properties in a subtropical gyre: implications for the evaluation of
interannual trends of biogeochemical properties. Journal of Geophysical
Research, 115(C9), C09012, doi:10.1029/2009JC005865.
Lehodey, P., M. Bertignac, J. Hampton, A. Lewis, and J. Picaut, 1997: El Niño Southern
Oscillation and tuna in the western Pacific. Nature, 389(6652), 715-718.
Lehodey, P., J. Alheit, M. Barange, T. Baumgartner, G. Beaugrand, K. Drinkwater, J.M.
Fromentin, S.R. Hare, G. Ottersen, R.I. Perry, C. Roy, C.D. van der Lingen, and
F. Werner, 2006: Climate variability, fish, and fisheries. Journal of Climate,
19(20), 5009-5030.
Lehodey, P., I. Senina, and R. Murtugudde, 2008: A spatial ecosystem and populations
dynamics model (SEAPODYM) – modeling of tuna and tuna-like populations.
Progress in Oceanography, 78(4), 304-318.
Lehodey, P., R. Murtugudde, and I. Senina, 2010a: Bridging the gap from ocean models
to population dynamics of large marine predators: a model of mid-trophic
functional groups. Progress in Oceanography, 84(1-2), 69-84.
Lehodey, P., I. Senina, J. Sibert, L. Bopp, B. Calmettes, J. Hampton, and R. Murtugudde,
2010b: Preliminary forecasts of Pacific bigeye tuna population trends under
the A2 IPCC scenario. Progress in Oceanography, 86, 302-315.
Lehodey, P., J. Hampton, R.W. Brill, S. Nicol, I. Senina, B. Calmettes, H.-O. rtner, L. Bopp,
T. Ilyina, J.D. Bell, and J. Sibert, 2011: Vulnerability of the ocean fisheries in the
tropical Pacific to climate change. In: Vulnerability of Tropical Pacific Fisheries and
Aquaculture to Climate Change [Bell, J.D., J.E. Johnson, and A.J. Hobday (eds.)].
Secretariat of the Pacific Community, Noumea, New Caledonia, pp. 433-492.
Lehodey, P., I. Senina, B. Calmettes, J. Hampton, and S. Nicol, 2013: Modelling the
impact of climate change on Pacific skipjack tuna population and fisheries.
Climatic Change, 119(1), 95-109.
Lejeusne, C., P. Chevaldonné, C. Pergent-Martini, C.F. Boudouresque, and T. Pérez,
2010: Climate change effects on a miniature ocean: the highly diverse, highly
impacted Mediterranean Sea. Trends in Ecology & Evolution, 25(4), 250-260.
Levin, L.A., W. Ekau, A.J. Gooday, F. Jorissen, J.J. Middelburg, S.W.A. Naqvi, C. Neira,
N.N. Rabalais, and J. Zhang, 2009: Effects of natural and human-induced
hypoxia on coastal benthos. Biogeosciences, 6(10), 2063-2098.
Levitus, S., J. Antonov, and T. Boyer, 2005: Warming of the world ocean, 1955-2003.
Geophysical Research Letters, 32(2), L02604, doi:10.1029/2004GL021592.
Levitus, S., J.I. Antonov, T.P. Boyer, R.A. Locarnini, H.E. Garcia, and A.V. Mishonov,
2009: Global ocean heat content 1955-2008 in light of recently revealed
instrumentation problems. Geophysical Research Letters, 36(7), L07608,
doi:10.1029/2008GL037155.
Lewis, E.L. and N.P. Fofonoff, 1979: A practical salinity scale. Journal of Physical
Oceanography, 9(2), 446.
Li, L., J. Xu, and R. Cai, 2002: Trends of sea level rise in the South China Sea during
the 1990s: an altimetry result. Chinese Science Bulletin, 47(7), 582-585.
Li, S., K.-F. Yu, T.-R. Chen, Q. Shi, and H.-L. Zhang, 2011: Assessment of coral bleaching
using symbiotic zooxanthellae density and satellite remote sensing data in the
Nansha Islands, South China Sea. Chinese Science Bulletin, 56(10), 1031-1037.
Lin, C., X. Ning, J. Su, Y. Lin, and B. Xu, 2005: Environmental changes and the
responses of the ecosystems of the Yellow Sea during 1976-2000. Journal of
Marine Systems, 55(3), 223-234.
Lin, G. and Q. Yang, 2011: Interdecadal variation of distribution features of
phytoplankton in the Taiwan Strait to global climate change. Chinese Journal
of Applied & Environmental Biology, 17(5), 615-623.
Lindegren, M., C. llmann, A. Nielsen, K. Brander, B.R. MacKenzie, and N.C.
Stenseth, 2010: Ecological forecasting under climate change: the case of Baltic
cod. Proceedings of the Royal Society B, 277(1691), 2121-2130.
Lindley, J.A. and R.R. Kirby, 2010: Climate-induced changes in the North Sea
Decapoda over the last 60 years. Climate Research, 42(3), 257-264.
Lindley, J.A., G. Beaugrand, C. Luczak, J.M. Dewarumez, and R.R. Kirby, 2010:
Warm-water decapods and the trophic amplification of climate in the North
Sea. Biology Letters, 6(6), 773-776.
Ling, S.D., 2008: Range expansion of a habitat-modifying species leads to loss of
taxonomic diversity: a new and impoverished reef state. Oecologia, 156(4),
883-894.
Ling, S.D., C.R. Johnson, S. Frusher, and C.K. King, 2008: Reproductive potential of a
marine ecosystem engineer at the edge of a newly expanded range. Global
Change Biology, 14(4), 907-915.
Ling, S.D., C.R. Johnson, S.D. Frusher, and K.R. Ridgway, 2009: Overfishing reduces
resilience of kelp beds to climate-driven catastrophic phase shift. Proceedings
of the National Academy of Sciences of the United States of America, 106(52),
22341-22345.
Link, J.S., B. Bogstad, H. Sparholt, and G.R. Lilly, 2009a: Trophic role of Atlantic cod
in the ecosystem. Fish and Fisheries, 10(1), 58-87.
Link, J.S., W.T. Stockhausen, G. Skaret, W. Overholtz, B.A. Megrey, H. Gjosaeter, S.
Gaichas, A. Dommasnes, J. Falk-Petersen, J. Kane, F.J. Mueter, K.D. Friedland,
and J.A. Hare, 2009b: A comparison of biological trends from four marine
ecosystems: synchronies, differences, and commonalities. Progress in
Oceanography, 81(1-4), 29-46.
Litzow, M.A. and L. Ciannelli, 2007: Oscillating trophic control induces community
reorganization in a marine ecosystem. Ecology Letters, 10(12), 1124-1134.
Litzow, M.A., J.D. Urban, and B.J. Laurel, 2008: Increased spatial variance accompanies
reorganization of two continental shelf ecosystems. Ecological Applications,
18(6), 1331-1337.
Liu, C., Y. Du, Q. Zhang, T. Chen, and D. Wang, 2007: Responses of oceans to global
warming and observation evidence in the South China Sea. Advances in
Climate Change Research, 3(1), 8-13.
Livingston, P.A. and S. Tjelmeland, 2000: Fisheries in boreal ecosystems. International
Council for the Exploration of the Sea (ICES) Journal of Marine Science, 57(3),
619-627.
Llewellyn, L.E., 2010: Revisiting the association between sea surface temperature
and the epidemiology of fish poisoning in the South Pacific: reassessing the
link between ciguatera and climate change. Toxicon, 56(5), 691-697.
Lo Iacono, C., M.A. Mateo, E. Gràcia, L. Guasch, R. Carbonell, L. Serrano, O. Serrano,
and J. Dañobeitia, 2008: Very high-resolution seismo-acoustic imaging of
seagrass meadows (Mediterranean Sea): implications for carbon sink estimates.
Geophysical Research Letters, 35, L18601, doi:10.1029/2008GL034773.
Longhurst, A., 1998. Ecological Geography of the Sea. Academic Press, San Diego,
CA, USA and London, UK, 398 pp.
Lotze, H.K., M. Coll, A.M. Magera, C. Ward-Paige, and L. Airoldi, 2011: Recovery of
marine animal populations and ecosystems. Trends in Ecology & Evolution,
26(11), 595-605.
Lough, J.M., 2012: Small change, big difference: sea surface temperature distributions
for tropical coral reef ecosystems, 1950-2011. Journal of Geophysical Research:
Oceans, 117(C9), C09018, doi:10.1029/2012JC008199.
Lough, J.M., G.A. Meehl, and M.J. Salinger, 2011: Observed in projected changes in
surface climate of the tropical Pacific. In: Vulnerability of Tropical Pacific
Fisheries and Aquaculture to Climate Change [Bell, J.D., J.E. Johnson, and A.J.
Hobday (eds.)]. Secretariat of the Pacific Community, Noumea, New Caledonia,
pp. 49-99.
1724
Chapter 30 The Ocean
30
Lucey, S.M. and J.A. Nye, 2010: Shifting species assemblages in the Northeast US
Continental Shelf Large Marine Ecosystem. Marine Ecology Progress Series,
415, 23-33.
Luchetta, A., C. Cantoni, and G. Catalano, 2010: New observations of CO
2
-induced
acidification in the northern Adriatic Sea over the last quarter century.
Chemistry and Ecology, 26(Suppl. 1), 1-17.
Ludwig, W., E. Dumont, M. Meybeck, and S. Heussner, 2009: River discharges of water
and nutrients to the Mediterranean and Black Sea: major drivers for ecosystem
changes during past and future decades? Progress in Oceanography, 80(3-4),
199-217.
Luisetti, T., E.L. Jackson, and R.K. Turner, 2013: Valuing the European ‘coastal blue
carbon’storage benefit. Marine Pollution Bulletin, 71(1-2), 101-106.
Lundsten, L., J.P. Barry, G.M. Cailliet, D.A. Clague, A.P. DeVogelaere, and J.B. Geller,
2009: Benthic invertebrate communities on three seamounts off southern and
central California, USA. Marine Ecology Progress Series, 374, 23-32.
Ma, Z., Z. Xu, and J. Zhou, 2009: Effect of global warming on the distribution of Lucifer
intermedius and L. hanseni (Decapoda) in the Changjiang estuary. Progress
in Natural Science, 19(10), 1389-1395.
Mabus, R., 2010: Department of the Navy’s Energy Program for Security and
Independence. United States Department of the Navy, DIANE Publishing,
Darby, PA, USA, 24 pp.
Mackas, D.L., R. Goldblatt, and A.G. Lewis, 1998: Interdecadal variation in
developmental timing of Neocalanus plumchrus populations at Ocean Station
P in the subarctic North Pacific. Canadian Journal of Fisheries and Aquatic
Sciences, 55(8), 1878-1893.
MacKenzie, B.R., and D. Schiedek, 2007a: Daily ocean monitoring since the 1860s
shows record warming of northern European seas. Global Change Biology,
13(7), 1335-1347.
MacKenzie, B.R., and D. Schiedek, 2007b: Long-term sea surface temperature
baselines-time series, spatial covariation and implications for biological
processes. Journal of Marine Systems, 68(3-4), 405-420.
MacKenzie, B.R., H. Gislason, C. Möllmann, and F.W. Köster, 2007: Impact of 21
st
century climate change on the Baltic Sea fish community and fisheries. Global
Change Biology, 13(7), 1348-1367.
MacKenzie, B.R., H.E.M. Meier, M. Lindegren, S. Neuenfeldt, M. Eero, T. Blenckner,
M.T. Tomczak, and S. Niiranen, 2012: Impact of climate change on fish population
dynamics in the Baltic Sea: a dynamical downscaling investigation. AMBIO: A
Journal of the Human Environment, 41(6), 626-636.
MacNeil, M.A., N.A. Graham, J.E. Cinner, N.K. Dulvy, P.A. Loring, S. Jennings, N.V.
Polunin, A.T. Fisk, and T.R. McClanahan, 2010: Transitional states in marine
fisheries: adapting to predicted global change. Philosophical Transactions of
the Royal Society B, 365(1558), 3753-3763.
Madsen, K.S. and N.K. Højerslev, 2009: Long-term temperature and salinity records from
the Baltic Sea transition zone. Boreal Environment Research, 14(1), 125-131.
Maier, C., J. Hegeman, M.G. Weinbauer, and J.-P. Gattuso, 2009: Calcification of the
cold-water coral Lophelia pertusa, under ambient and reduced pH.
Biogeosciences, 6(8), 1671-1680.
Maier, C., A. Schubert, M.M.B. nchez, M.G. Weinbauer, P. Watremez, and J.-P.
Gattuso, 2013: End of the century pCO
2
levels do not impact calcification in
Mediterranean cold-water corals. PLoS ONE, 8(4), e62655, doi:10.1371/
journal.pone.0062655.
Makino, M., H. Matsuda, and Y. Sakurai, 2009: Expanding fisheries co-management
to ecosystem-based management: a case in the Shiretoko World Natural
Heritage area, Japan. Marine Policy, 33(2), 207-214.
Manhique, A.J., C.J.C. Reason, L. Rydberg, and N. Fauchereau, 2011: ENSO and Indian
Ocean sea surface temperatures and their relationships with tropical temperate
troughs over Mozambique and the Southwest Indian Ocean. International
Journal of Climatology, 31(1), 1-13.
Mantua, N.J. and S.R. Hare, 2002: The Pacific Decadal Oscillation. Journal of
Oceanography, 58(1), 35-44.
Mantua, N.J., S.R. Hare, Y. Zhang, J.M. Wallace, and R.C. Francis, 1997: A Pacific
interdecadal climate oscillation with impacts on salmon production. Bulletin
of the American Meteorological Society, 78(6), 1069-1079.
Manzello, D.P., 2010: Coral growth with thermal stress and ocean acidification:
lessons from the eastern tropical Pacific. Coral Reefs, 29(3), 749-758.
Manzello, D.P., J.A. Kleypas, D.A. Budd, C.M. Eakin, P.W. Glynn, and C. Langdon, 2008:
Poorly cemented coral reefs of the eastern tropical Pacific: possible insights
into reef development in a high-CO
2
world. Proceedings of the National
Academy of Sciences of the United States of America, 105(30), 10450-10455.
Marbà, N. and C.M. Duarte, 2010: Mediterranean warming triggers seagrass
(Posidonia oceanica) shoot mortality. Global Change Biology, 16(8), 2366-
2375.
Martin, J.H. and S.E. Fitzwater, 1988: Iron deficiency limits phytoplankton growth in
the north-east Pacific subarctic. Nature, 331(6154), 341-343.
Martin, S. and J.-P. Gattuso, 2009: Response of Mediterranean coralline algae to
ocean acidification and elevated temperature. Global Change Biology, 15(8),
2089-2100.
Marullo, S., V. Artale, and R. Santoleri, 2011: The SST multi-decadal variability in the
Atlantic-Mediterranean region and its relation to AMO. Journal of Climate,
24(16), 4385-4401.
Massuti, E., S. Monserrat, P. Oliver, J. Moranta, J.L. pez-Jurado, M. Marcos, M. Hidalgo,
B. Guijarro, A. Carbonell, and P. Pereda, 2008: The influence of oceanographic
scenarios on the population dynamics of demersal resources in the western
Mediterranean: hypothesis for hake and red shrimp off Balearic Islands.
Journal of Marine Systems, 71(3-4), 421-438.
Matsuno, T., J.S. Lee, and S. Yanao, 2009: The Kuroshio exchange with the South and
East China Seas. Ocean Science, 5(3), 303-312.
McAllister, C.D., T.R. Parsons, and J.D.H. Strickland, 1960: Primary productivity and
fertility at Station “P” in the North-East Pacific Ocean. International Council
for the Exploration of the Sea (ICES) Journal of Marine Science, 25(3), 240-
259.
McCarthy, J.J., C. Garside, J.L. Nevins, and R.T. Barber, 1996: New production along
140°W in the equatorial Pacific during and following the 1992 El Niño event.
Deep-Sea Research Part II: Topical Studies in Oceanography, 43(4-6), 1065-
1093.
McClain, C.R., S.R. Signorini, and J.R. Christian, 2004: Subtropical gyre variability
observed by ocean-color satellites. Deep-Sea Research Part II: Topical Studies
in Oceanography, 51(1-3), 281-301.
McClanahan, T.R., M. Ateweberhan, N.A.J. Graham, S.K. Wilson, C.R. Sebastián,
M.M.M. Guillaume, and J.H. Bruggemann, 2007: Western Indian Ocean coral
communities: bleaching responses and susceptibility to extinction. Marine
Ecology Progress Series, 337, 1-13.
McClanahan, T.R., J.E. Cinner, J. Maina, N.A.J. Graham, T.M. Daw, S.M. Stead, A.
Wamukota, K. Brown, M. Ateweberhan, V. Venus, and N.V.C. Polunin, 2008:
Conservation action in a changing climate. Conservation Letters, 1(2), 53-59.
McClanahan, T.R., M. Ateweberhan, J. Omukoto, and L. Pearson, 2009a: Recent
seawater temperature histories, status, and predictions for Madagascar’s coral
reefs. Marine Ecology Progress Series, 380, 117-128.
McClanahan, T.R., E. Weil, and J. Maina, 2009b: Strong relationship between coral
bleaching and growth anomalies in massive Porites. Global Change Biology,
15(7), 1804-1816.
McClanahan, T.R., E. Weil, J. Cortés, A.H. Baird, and M. Ateweberhan, 2009c:
Consequences of coral bleaching for sessile reef organisms. In: Coral
Bleaching: Patterns, Processes, Causes and Consequences, [van Oppen, M.J.H.
and J.M. Lough (eds.)]. Springer-Verlag, Berlin, Germany, pp. 121-138.
McClatchie, S., R. Goericke, R. Cosgrove, G. Auad, and R. Vetter, 2010: Oxygen in the
Southern California Bight: multidecadal trends and implications for demersal
fisheries. Geophysical Research Letters, 37(19), L19602, doi:10.1029/
2010GL044497.
McGowan, J.A., D.R. Cayan, and L.M. Dorman, 1998: Climate-ocean variability and
ecosystem response in the Northeast Pacific. Science, 281(5374), 210-216.
McIlgorm, A., S. Hanna, G. Knapp, P. Le Floc’H, F. Millerd, and M. Pan, 2010: How
will climate change alter fishery governance? Insights from seven international
case studies. Marine Policy, 34(1), 170-177.
McKinnell, S.M. and M.J. Dagg (eds.), 2010: Marine Ecosystems of the North Pacific
Ocean, 2003-2008. PICES Special Publication No. 4, North Pacific Marine
Science Organization (PICES), Sydney, BC, Canada, 393 pp.
McKinnell, S.M., S. Batten, S.J. Bograd, J.L. Boldt, N. Bond, S. Chiba, M.J. Dagg, M.G.G.
Foreman, G.L. Hunt Jr., J.R. Irvine, O.N. Katugin, V. Lobanov, D.L. Mackas, P.
Mundy, V. Radchenko, Y.J. Ro, H. Sugisaki, F.A. Whitney, A. Yatsu, and S. Yoo,
2010: Status and trends of the North Pacific Ocean, 2003-2008. In: Marine
Ecosystems of the North Pacific Ocean, 2003-2008 [McKinnell, S.M. and M.J.
Dagg (eds.)]. PICES Special Publication No. 4, North Pacific Marine Science
Organization (PICES), Sydney, BC, Canada, pp. 1-55.
McLeod, E., R. Moffitt, A. Timmermann, R. Salm, L. Menviel, M.J. Palmer, E.R. Selig,
K.S. Casey, and J.F. Bruno, 2010: Warming seas in the Coral Triangle: coral reef
vulnerability and management implications. Coastal Management, 38(5), 518-
539.
1725
The Ocean Chapter 30
30
McLeod, E., G.L. Chmura, S. Bouillon, R. Salm, M. Björk, C.M. Duarte, C.E. Lovelock,
W.H. Schlesinger, and B.R. Silliman, 2011: A blueprint for blue carbon: toward
an improved understanding of the role of vegetated coastal habitats in
sequestering CO
2
. Frontiers in Ecology and the Environment, 9(10), 552-560.
McMichael, A.J., R.E. Woodruff, and S. Hales, 2006: Climate change and human
health: present and future risks. Lancet, 367(9513), 859-869.
McNeil, B.I. and R.J. Matear, 2008: Southern Ocean acidification: a tipping point at
450-ppm atmospheric CO
2
. Proceedings of the National Academy of Sciences
of the United States of America, 105(48), 18860-18864.
McPhaden, M.J. and D. Zhang, 2002: Slowdown of the meridional overturning
circulation in the upper Pacific Ocean. Nature, 415(6872), 603-608.
MEA, 2005: Ecosystems and Human Well-Being: Current State and Trends: Findings
of the Condition and Trends Working Group, Vol. 1 [Hassan, R., R. Scholes, and
N. Ash (eds.)]. Millennium Ecosystem Assessment (MEA), Island Press,
Washington, DC, USA, 917 pp.
Meehl, G.A. and J.M. Arblaster, 2011: Decadal variability of Asian–Australian
Monsoon–ENSO–TBO relationships. Journal of Climate, 24(18), 4925-4940.
Meehl, G.A., A. Hu, and B.D. Santer, 2009: The mid-1970s climate shift in the Pacific
and the relative roles of forced versus inherent decadal variability. Journal of
Climate, 22(3), 780-792.
Menard, F., F. Marsac, E. Bellier, and B. Cazelles, 2007: Climatic oscillations and tuna
catch rates in the Indian Ocean: a wavelet approach to time series analysis.
Fisheries Oceanography, 16(1), 95-104.
Mendelssohn, R. and F. Schwing, 2002: Common and uncommon trends in SST and
wind stress in the California and Peru–Chile current systems. Progress in
Oceanography, 53(2), 141-162.
Menge, B.A., F. Chan, K.J. Nielsen, E. di Lorenzo, and J. Lubchenco, 2009: Climatic
variation alters supply-side ecology: impact of climate patterns on phytoplankton
and mussel recruitment. Ecological Monographs, 79(3), 379-395.
Merino, G., M. Barange, C. Mullon, and L. Rodwell, 2010: Impacts of global
environmental change and aquaculture expansion on marine ecosystems.
Global Environmental Change: Human and Policy Dimensions, 20(4), 586-596.
Merino, G., M. Barange, J.L. Blanchard, J. Harle, R. Holmes, I. Allen, E.H. Allison, M.C.
Badjeck, N.K. Dulvy, J. Holt, S. Jennings, C. Mullon, and L. Rodwell, 2012: Can
marine fisheries and aquaculture meet fish demand from a growing human
population in a changing climate? Global Environmental Change, 22(4), 795-
806.
Mertz, O., K. Halsnæs, J.E. Olesen, and K. Rasmussen, 2009: Adaptation to climate
change in developing countries. Environmental Management, 43(5), 743-752.
Mestas-Nuñez, A.M. and A.J. Miller, 2006: Interdecadal variability and climate
change in the eastern tropical Pacific: a review. Progress in Oceanography,
69(2-4), 267-284.
Mieszkowska, N., S.J. Hawkins, M.T. Burrows, and M.A. Kendall, 2007: Long-term
changes in the geographic distribution and population structures of Osilinus
lineatus (Gastropoda: Trochidae) in Britain and Ireland. Journal of the Marine
Biological Association of the United Kingdom, 87(2), 537-545.
Miller, K.A., G.R. Munro, U.R. Sumaila, and W.W.L. Cheung, 2013: Governing marine
fisheries in a changing climate: a game-theoretic perspective. Canadian Journal
of Agricultural Economics, 61(2 SI), 309-334.
Mills, D.J., L. Westlund, G. Graaf, Y. de Kura, R. Willman, and K. Kelleher, 2011: Under-
reported and undervalued: small-scale fisheries in the developing world. In:
Small-Scale Fisheries Management: Frameworks and Approaches for the
Developing World [Pomeroy, R.S. and N.L. Andrew (eds.)]. CABI (formerly CAB
International), Wallingford, UK. pp. 1-15.
Minobe, S., 1997: A 50-70 year climatic oscillation over the North Pacific and North
America. Geophysical Research Letters, 24(6), 683-686.
Mohan, R., K. Verma, L.P. Mergulhao, D.K. Sinha, S. Shanvas, and M.V.S. Guptha, 2006:
Seasonal variation of pteropods from the Western Arabian Sea sediment trap.
Geo-Marine Letters, 26(5), 265-273.
Molinero, J.C., F. Ibanez, S. Souissi, M. Chifflet, and P. Nival, 2005: Phenological
changes in the Northwestern Mediterranean copepods Centropages typicus
and Temora stylifera linked to climate forcing. Oecologia, 145(4), 640-649.
Möllmann, C., G. Kornilovs, M. Fetter, F.W. Köster, and H.H. Hinrichsen, 2003: The
marine copepod, Pseudocalanus elongatus, as a mediator between climate
variability and fisheries in the Central Baltic Sea. Fisheries Oceanography,
12(4-5), 360-368.
Möllmann, C., G. Kornilovs, M. Fetter, and F.W. Köster, 2005: Climate, zooplankton,
and pelagic fish growth in the central Baltic Sea. International Council for the
Exploration of the Sea (ICES) Journal of Marine Science, 62(7), 1270-1280.
Möllmann, C., R. Diekman, B. Müller-Karulis, G. Kornilovs, M. Plikshs, and P. Axe,
2009: Reorganization of a large marine ecosystem due to atmospheric and
anthropogenic pressure: a discontinuous regime shift in the Central Baltic Sea.
Global Change Biology, 15(6), 1377-1393.
Monteiro, P.M.S., A.K. van der Plas, J.L. Mélice, and P. Florenchie, 2008: Interannual
hypoxia variability in a coastal upwelling system: ocean-shelf exchange, climate
and ecosystem-state implications. Deep Sea Research Part I: Oceanographic
Research Papers, 55(4), 435-450.
Morita, K., S.H. Morita, and M.A. Fukuwaka, 2006: Population dynamics of Japanese
pink salmon (Oncorhynchus gorbuscha): are recent increases explained by
hatchery programs or climatic variations? Canadian Journal of Fisheries and
Aquatic Sciences, 63(1), 55-62.
Moron, V., 2003: Long-term variability of the Mediterranean Sea surface temperature
(1856-2000). Compte Rendus Geoscience, 335(9), 721-727.
Mountain, D.G. and J. Kane, 2010: Major changes in the Georges Bank ecosystem,
1980s to the 1990s. Marine Ecology Progress Series, 398, 81-91.
Movilla, J., E. Calvo, C. Pelejero, R. Coma, E. Serrano, P. Fernández-Vallejo, and M.
Ribes, 2012: Calcification reduction and recovery in native and non-native
Mediterranean corals in response to ocean acidification. Journal of
Experimental Marine Biology and Ecology, 438, 144-153.
Moy, A.D., W.R. Howard, S.G. Bray, and T.W. Trull, 2009: Reduced calcification in modern
Southern Ocean planktonic foraminifera. Nature Geoscience, 2(4), 276-280.
Mueter, F.J. and M.A. Litzow, 2008: Sea ice retreat alters the biogeography of the
Bering Sea continental shelf. Ecological Applications, 18(2), 309-320.
Muhling, B.A., S.-K. Lee, J.T. Lamkin, and Y. Liu, 2011: Predicting the effects of climate
change on bluefin tuna (Thunnus thynnus) spawning habitat in the Gulf of
Mexico. International Council for the Exploration of the Sea (ICES) Journal of
Marine Science, 68(6), 1051-1062.
Mumby, P.J., J.R.M. Chisholm, A.J. Edwards, C.D. Clark, E.B. Roark, S. Andréfouët, and
J. Jaubert, 2001: Unprecedented bleaching-induced mortality in Porites spp.
at Rangiroa Atoll, French Polynesia. Marine Biology, 139(1), 183-189.
Mumby, P.J., R. Vitolo, and D.B. Stephenson, 2011: Temporal clustering of tropical
cyclones and its ecosystem impacts. Proceedings of the National Academy of
Sciences of the United States of America, 108(43), 17626-17630.
Munday, P.L., D.L. Dixson, J.M. Donelson, G.P. Jones, M.S. Pratchett, G.V. Devitsina,
and K.B. Doving, 2009: Ocean acidification impairs olfactory discrimination
and homing ability of a marine fish. Proceedings of the National Academy of
Sciences of the United States of America, 106(6), 1848-1852.
Murray, J.W., H.W. Jannasch, S. Honjo, R.F. Anderson, W.S. Reeburgh, Z. Top, G.E.
Friederich, L.A. Codispoti, and E. Izdar, 1989: Unexpected changes in the
oxic/anoxic interface in the Black Sea. Nature, 338(6214), 411-413.
Myers, S.S. and J.A. Patz, 2009: Emerging threats to human health from global
environmental change. Annual Review of Environment and Resources, 34,
223-252.
Nakamura, N., H. Kayanne, H. Iijima, T.R. McClanahan, S.K. Behera, and T. Yamagata,
2011: Footprints of IOD and ENSO in the Kenyan coral record. Geophysical
Research Letters, 38(24), L24708, doi:10.1029/2011GL049877.
Nakanowatari, T., K.I. Ohshima, and M. Wakatsuchi, 2007: Warming and oxygen
decrease of intermediate water in the northwestern North Pacific, originating
from the Sea of Okhotsk, 1955-2004. Geophysical Research Letters, 34(4),
L04602, doi:10.1029/2006GL028243.
Narayan, N., A. Paul, S. Mulitza, and M. Schulz, 2010: Trends in coastal upwelling
intensity during the late 20
th
century. Ocean Science, 6(3), 815-823.
Naylor, R.L., R.W. Hardy, D.P. Bureau, A. Chiu, M. Elliott, A.P. Farrell, I. Forster, D.M.
Gatlin, R.J. Goldburg, K. Hua, and P.D. Nichols, 2009: Feeding aquaculture in
an era of finite resources. Proceedings of the National Academy of Sciences
of the United States of America, 106(42), 15103-15110.
Nellemann, C., E. Corcoran, C.M. Duarte, L. Valdés, C. De Young, L. Fonseca, and G.
Grimsditch (eds.), 2009: Blue Carbon: The Role of Healthy Oceans in Binding
Carbon – A Rapid Response Assessment. Report produced as an Inter-agency
collaboration between the United Nations Environment Programme (UNEP),
the Food and Agricultural Organization of the United Nations (FAO) and the
Intergovernmental Oceanographic Commission (IOC)/ United Nations
Educational, Scientific and Cultural Organization (UNESCO), UNEP, Nairobi,
Kenya and Global Resource Information Database (GRID)-Arendal, Arendal,
Norway, 78 pp.
Nicol, S.J., V. Allain, G.M. Pilling, J. Polovina, M. Coll, J. Bell, P. Dalzell, P. Sharples, R.
Olson, S. Griffiths, J.M. Dambacher, J. Young, A. Lewis, J. Hampton, J. Jurado
Molina, S. Hoyle, K. Briand, N. Bax, P. Lehodey, and P. Williams, 2013: An ocean
1726
Chapter 30 The Ocean
30
observation system for monitoring the affects of climate change on the
ecology and sustainability of pelagic fisheries in the Pacific Ocean. Climatic
Change, 119(1), 131-145.
Ning, X., C. Lin, Q. Hao, C. Liu, F. Le, and J. Shi, 2009: Long term changes in the
ecosystem in the northern South China Sea during 1976-2004. Biogeosciences,
6(10), 2227-2243.
Ning, X., C. Lin, J. Su, C. Liu, Q. Hao, and F. Le, 2011: Long-term changes of dissolved
oxygen, hypoxia, and the responses of the ecosystems in the East China Sea
from 1975 to 1995. Journal of Oceanography, 67(1), 59-75.
Norman-López, A., É. Plagányi, T. Skewes, E. Poloczanska, D. Dennis, M.T. Gibbs, and P.
Bayliss, 2013: Linking physiological, population and socio-economic assessments
of climate-change impacts on fisheries. Fisheries Research, 148, 18-26.
Nøttestad, L., K.R. Utne, V. Anthonypillai, Ø. Tangen, J.W. Valdemarsen, G.J. Óskarsson,
S. Sveinbjörnsoon, S. Jónsson, H. Debes, E. Mortensen, L. Smith, A. Ólafsdottir,
and J.A. Jacobsen, 2013: Cruise Report from the Coordinated Ecosystem Survey
(IESSMS) with M/V “Libas”, M/V “Eros”, M/V “Finnur Fríiand R/V “Arni
Fridriksson” in the Norwegian Sea and Surrounding Waters, 2 July - 9 August
2013.Working Document to ICES Working Group on Widely Distributed Stocks
(WGWIDE), 27 August - 2 September 2013, International Council for the
Exploration of the Sea (ICES) Headquarters, Copenhagen, Denmark, 42 pp.
Nye, J.A., J.S. Link, J.A. Hare, and W.J. Overholtz, 2009: Changing spatial distribution
of fish stocks in relation to climate and population size on the Northeast
United States continental shelf. Marine Ecology Progress Series, 393, 111-
129.
Obura, D. and S. Mangubhai, 2011: Coral mortality associated with thermal fluctuations
in the Phoenix Islands, 2002-2005. Coral Reefs, 30(3), 607-619.
Oguz, T. and V. Velikova, 2010: Abrupt transition of the northwestern Black Sea shelf
ecosystem from a eutrophic to an alternative pristine state. Marine Ecology
Progress Series, 405, 231-242.
Oguz, T., T. Cokacar, P. Malanotte-Rizzoli, and H.W. Ducklow, 2003: Climatic warming
and accompanying changes in the ecological regime of the Black Sea during
1990s. Global Biogeochemical Cycles, 17(3), 1088, doi:10.1029/2003GB002031.
Oguz, T., B. Fach, and B. Salihoglu, 2008: Invasion dynamics of the alien ctenophore
Mnemiopsis leidyi and its impact on anchovy collapse in the Black Sea. Journal
of Plankton Research, 30(12), 1385-1397.
Ohmura, A., 2009: Observed decadal variations in surface solar radiation and their
causes. Journal of Geophysical Research: Atmospheres, 114(D10), D00D05,
doi:10.1029/2008JD011290.
Orr, J.C., V.J. Fabry, O. Aumont, L. Bopp, S.C. Doney, R.A. Feely, A. Gnanadesikan, N.
Gruber, A. Ishida, F. Joos, R.M. Key, K. Lindsay, E. Maier-Reimer, R. Matear, P.
Monfray, A. Mouchet, R.G. Najjar, G.K. Plattner, K.B. Rodgers, C.L. Sabine, J.L.
Sarmiento, R. Schlitzer, R.D. Slater, I.J. Totterdell, M.F. Weirig, Y. Yamanaka, and
A. Yool, 2005: Anthropogenic ocean acidification over the twenty-first century
and its impact on calcifying organisms. Nature, 437(7059), 681-686.
Oude Elferink, A.G., 2012: Governance principles for areas beyond national jurisdiction.
The International Journal of Marine and Coastal Law, 27(2), 205-259.
Overland, J., S. Rodionov, S. Minobe, and N. Bond, 2008: North Pacific regime shifts:
definitions, issues and recent transitions. Progress in Oceanography, 77(2-3),
92-102.
Overland, J.E., J. Alheit, A. Bakun, J.W. Hurrell, D.L. Mackas, and A.J. Miller, 2010:
Climate controls on marine ecosystems and fish populations. Journal of Marine
Systems, 79(3-4), 305-315.
Parmesan, C., C. Duarte, E. Poloczanska, A.J. Richardson, and M.C. Singer, 2011:
Commentary: overstretching attribution. Nature Climate Change, 1(1), 2-4.
Parrish, R.H., F.B. Schwing, and R. Mendelssohn, 2000: Mid-latitude wind stress: the
energy source for climatic shifts in the North Pacific Ocean. Fisheries
Oceanography, 9(3), 224-238.
Pascual, M., X. Rodó, S.P. Ellner, R. Colwell, and M.J. Bouma, 2000: Cholera dynamics
and El Niño-Southern Oscillation. Science, 289(5485), 1766-1769.
Patz, J.A., D. Campbell-Lendrum, T. Holloway, and J.A. Foley, 2005: Impact of regional
climate change on human health. Nature, 438(7066), 310-317.
Pauly, D. and V. Christensen, 1995: Primary production required to sustain global
fisheries. Nature, 374(6519), 255-257.
Pecl, G., S. Frusher, C. Gardner, M. Haward, A. Hobday, S. Jennings, M. Nursey-Bray,
A. Punt, H. Revill, and I. van Putten, 2009: The East Coast Tasmanian Rock
Lobster Fishery Vulnerability to Climate Change Impacts and Adaptation
Response Options. Report to the Australian Government Department of
Climate Change, Australia, Commonwealth of Australia, Canberra, ACT, Australia,
114 pp.
Peñaflor, E., W. Skirving, A. Strong, S. Heron, and L. David, 2009: Sea-surface
temperature and thermal stress in the Coral Triangle over the past two
decades. Coral Reefs, 28(4), 841-850.
Pendleton, L., D.C. Donato, B.C. Murray, S. Crooks, W.A. Jenkins, S. Sifleet, C. Craft,
J.W. Fourqurean, J.B. Kauffman, N. Marbà, P. Megonigal, E. Pidgeon, D. Herr, D.
Gordon, and A. Baldera, 2012: Estimating global “blue carbon” emissions
from conversion and degradation of vegetated coastal ecosystems. PLoS ONE,
7(9), e43542, doi:10.1371/journal.pone.004354.
Pennington, J.T., K.L. Mahoney, V.S. Kuwahara, D.D. Kolber, R. Calienes, and F.P.
Chavez, 2006: Primary production in the eastern tropical Pacific: a review.
Progress in Oceanography, 69(2-4), 285-317.
Perry, A.L., P.J. Low, J.R. Ellis, and J.D. Reynolds, 2005: Climate change and distribution
shifts in marine fishes. Science, 308(5730), 1912-1915.
Perry, R.I., P. Cury, K. Brander, S. Jennings, C. Möllmann, and B. Planque, 2010:
Sensitivity of marine systems to climate and fishing: concepts, issues and
management responses. Journal of Marine Systems, 79, 427-435.
Peterson, W.T. and F.B. Schwing, 2003: A new climate regime in northeast Pacific
ecosystems. Geophysical Research Letters, 30(17), 1896, doi:10.1029/
2003GL017528.
Pickering, T., B. Ponia, C. Hair, P.C. Southgate, E. Poloczanska, L. Della Patrona, A.
Teitelbaum, C.V. Mohan, M. Phillips, and J.D. Bell, 2011: Vulnerability of
aquaculture in the tropical Pacific to climate change. In: Vulnerability of
Tropical Pacific Fisheries and Aquaculture to Climate Change [Bell, J.D., J.E.
Johnson, and A.J. Hobday (eds.)]. Secretariat of the Pacific Community,
Noumea, New Caledonia, pp. 647-732.
Pierce, D.W., P.J. Gleckler, T.P. Barnett, B.D. Santer, and P.J. Durack, 2012: The fingerprint
of human-induced changes in the ocean’s salinity and temperature fields.
Geophysical Research Letters, 39(21), L21704, doi:10.1029/2012GL053389.
Pinnegar, J., T. Watt, and K. Kennedy, 2012: Climate Change Risk Assessment for the
Marine and Fisheries Sector. CCRA UK 2012 Climate Change Risk Assessment,
DEFRA Project Code GA0204, Department for Environment Food and Rural
Affairs (DEFRA), London, UK, 204 pp.
Pinsky, M.L., B. Worm, M.J. Fogarty, J.L. Sarmiento, and S.A. Levin, 2013: Marine taxa
track local climate velocities. Science, 341(6151), 1239-1242.
Pitcher, T.J., D. Kalikoski, K. Short, D. Varkey, and G. Pramod, 2009: An evaluation of
progress in implementing ecosystem-based management of fisheries in 33
countries. Marine Policy, 33(2), 223-232.
Pitt, N.R., E.S. Poloczanska, and A.J. Hobday, 2010: Climate-driven range changes in
Tasmanian intertidal fauna. Marine and Freshwater Research, 61(9), 963-970.
Plagányi, É.E., T. Skewes, M. Haddon, and N. Dowling, 2013: Risk management tools
for sustainable fisheries management under changing climate: a sea cucumber
example. Climatic Change, 119(1), 181-197.
Poloczanska, E.S., S.J. Hawkins, A.J. Southward, and M.T. Burrows, 2008: Modeling
the response of populations of competing species to climate change. Ecology,
89(11), 3138-3149.
Poloczanska, E.S., C.J. Brown, W.J. Sydeman, W. Kiessling, D.S. Schoeman, P.J. Moore,
K. Brander, J.F. Bruno, L.B. Buckley, M.T. Burrows, C.M. Duarte, B.S. Halpern, J.
Holding, C.V. Kappel, M.I. O’Connor, J.M. Pandolfi, C. Parmesan, F.B. Schwing,
S.A. Thompson, and A.J. Richardson, 2013: Global imprint of climate change
on marine life. Nature Climate Change, 3, 919-925.
Polovina, J.J. and P.A. Woodworth, 2012: Declines in phytoplankton cell size in the
subtropical oceans estimated from satellite remotely-sensed temperature and
chlorophyll, 1998-2007. Deep Sea Research Part II: Topical Studies in
Oceanography, 77-80(SI), 82-88.
Polovina, J.J., E.A. Howell, and M. Abecassis, 2008: Ocean’s least productive waters
are expanding. Geophysical Research Letters, 35(3), L03618, doi:10.1029/
2007GL031745.
Polovina, J.J., J.P. Dunne, P.A. Woodworth, and E.A. Howell, 2011: Projected expansion
of the subtropical biome and contraction of the temperate and equatorial
upwelling biomes in the North Pacific under global warming. International
Council for the Exploration of the Sea (ICES) Journal of Marine Science, 68(6),
986-995.
Pomeroy, R.S. and N.L. Andrew, 2011: Small-Scale Fisheries Management:
Frameworks and Approaches for the Developing World. CABI (formerly CAB
International), Wallingford, UK and Cambridge, MA, USA, 247 pp.
Pörtner, H.-O., 2002: Climate variations and the physiological basis of temperature
dependent biogeography: systemic to molecular hierarchy of thermal tolerance
in animals. Comparative Biochemistry and Physiology Part A: Molecular &
Integrative Physiology, 132(4), 739-761.
1727
The Ocean Chapter 30
30
Pörtner, H.-O., 2010: Oxygen- and capacity-limitation of thermal tolerance: a matrix
for integrating climate-related stressor effects in marine ecosystems. Journal
of Experimental Biology, 213(6), 881-893.
Pörtner, H.-O., B. Berdal, R. Blust, O. Brix, A. Colosimo, B. De Wachter, A. Giuliani, T.
Johansen, T. Fischer, R. Knust, G. Lannig, G. Naevdal, A. Nedenes, G. Nyhammer,
F.J. Sartoris, I. Serendero, P. Sirabella, S. Thorkildsen, and M. Zakhartsev, 2001:
Climate induced temperature effects on growth performance, fecundity and
recruitment in marine fish: developing a hypothesis for cause and effect
relationships in Atlantic cod (Gadus morhua) and common eelpout (Zoarces
viviparus). Continental Shelf Research, 21(18-19), 1975-1997.
Pörtner, H.-O., C. Bock, K. Deigweiher, and M. Lucassen, 2007: Environmental change
and the role of acid-base regulation in marine ectotherms. (Paper presented at
the Seventh International Congress of Comparative Physiology and Biochemistry,
Salvador, Bahia, Brazil, August 12-16, 2007). Comparative Biochemistry and
Physiology Part A: Molecular & Integrative Physiology, 148(Suppl. 1), S142,
www.sciencedirect.com/science/article/pii/S1095643307014195.
Power, S.B. and I.N. Smith, 2007: Weakening of the Walker Circulation and apparent
dominance of El Niño both reach record levels, but has ENSO really changed?
Geophysical Research Letters, 34(18), L18702, doi:10.1029/2007GL030854.
Power, S., T. Casey, C. Folland, A. Colman, and V. Mehta, 1999: Inter-decadal modulation
of the impact of ENSO on Australia. Climate Dynamics, 15(5), 319-324.
Prakash, S. and R. Ramesh, 2007: Is the Arabian Sea getting more productive?
Current Science, 92(5), 667-670.
Pratchett, M.S., P.L. Munday, S.K. Wilson, N.A.J. Graham, J.E. Cinner, D.R. Bellwood, G.P.
Jones, N.V.C. Polunin, and T.R. McClanahan, 2008: Effects of climate-induced
coral bleaching on coral-reef fishes – ecological and economic consequences.
Oceanography and Marine Biology: An Annual Review, 46, 251-296.
Pratchett, M.S., A.S. Hoey, S.K. Wilson, V. Messmer, and N.A.J. Graham, 2011a:
Changes in biodiversity and functioning of reef fish assemblages following
coral bleaching and coral loss. Diversity, 3(3), 424-452.
Pratchett, M.S., P.L. Munday, N.A.J. Graham, M. Kronen, S. Pinca, K. Friedman, T.D.
Brewer, J.D. Bell, S.K. Wilson, J.E. Cinner, J.P. Kinch, R.J. Lawton, A.J. Williams,
L. Chapman, F. Magron, and A. Webb, 2011b: Vulnerability of coastal fisheries
in the tropical Pacific to climate change. In: Vulnerability of Tropical Pacific
Fisheries and Aquaculture to Climate Change [Bell, J.D., J.E. Johnson, and A.J.
Hobday (eds.)]. Secretariat of the Pacific Community, Noumea, New Caledonia.
pp. 167-185.
Przeslawski, R., I. Falkner, M.B. Ashcroft, and P. Hutchings, 2012: Using rigorous
selection criteria to investigate marine range shifts. Estuarine, Coastal and
Shelf Science, 113, 205-212.
Qi, Q., R. Cai, and Q. Zhang, 2010: Low-frequency variability of the heat transport in
source area of Kuroshio and its relation to SST in China seas. Journal of
Oceanography in Taiwan Strait, 29(1), 106-113 (in Chinese).
Qiu, Y., Z. Lin, and Y. Wang, 2010: Responses of fish production to fishing and climate
variability in the northern South China Sea. Progress in Oceanography, 85(3-4),
197-212.
Rabalais, N.N., R.E. Turner, R.J. Díaz, and D. Justić, 2009: Global change and
eutrophication of coastal waters. International Council for the Exploration of
the Sea (ICES) Journal of Marine Science, 66(7), 1528-1537.
Rabalais, N.N., R.J. Diaz, L.A. Levin, R.E. Turner, D. Gilbert, and J. Zhang, 2010:
Dynamics and distribution of natural and human-caused hypoxia.
Biogeosciences, 7(2), 585-619.
Racault, M.F., C. Le Quéré, E. Buitenhuis, S. Sathyendranath, and T. Platt, 2012:
Phytoplankton phenology in the global ocean. Ecological Indicators, 14(1),
152-163.
Rahman, C., 2012: The implications of climate change for maritime security forces.
In: Climate Change and the Oceans: Gauging the Legal and Policy Currents in
the Asia Pacific and Beyond [Warner, R. and C. Schofield (eds.)]. Edward Elgar
Publishing, Cheltenham, UK and Northampton, MA, USA, pp. 167-199.
Raitsos, D.E., G. Beaugrand, D. Georgopoulos, A. Zenetos, A.M. Pancucci-
Papadopoulou, A. Theocharis, and E. Papathanassiou, 2010: Global climate
change amplifies the entry of tropical species into the Eastern Mediterranean
Sea. Limnology and Oceanography, 55(4), 1478-1484.
Raitsos, D.E., I. Hoteit, P.K. Prihartato, T. Chronis, G. Triantafyllou, and Y. Abualnaja,
2011: Abrupt warming of the Red Sea. Geophysical Research Letters, 38,
L14601, doi:10.1029/2011GL047984.
Randolph, M.F., C. Gaudin, S.M. Gourvenec, D.J. White, N. Boylan, and M.J. Cassidy,
2010: Recent advances in offshore geotechnics for deep water oil and gas
developments. Ocean Engineering, 38(7), 818-834.
Raven, J., K. Caldeira, H. Elderfield, O. Hoegh-Guldberg, P. Liss, U. Riebesell, J.
Shepherd, C. Turley, and A. Watson, 2005: Ocean Acidification due to Increasing
Atmospheric Carbon Dioxide. Policy document 12/05, The Royal Society,
London, UK, 60 pp.
Rayfuse, R. and R. Warner, 2012: Climate change mitigation activities in the ocean:
turning up the regulatory heat. In: Climate Change and the Oceans: Gauging
the Legal and Policy Currents in the Asia Pacific and Beyond [Warner, R. and C.
Schofield (eds.)]. Edward Elgar Publishing, Cheltenham, UK and Northampton,
MA, USA, pp. 234-262.
Rayner, N.A., D.E. Parker, E.B. Horton, C.K. Folland, L.V. Alexander, D.P. Rowell, E.C.
Kent, and A. Kaplan, 2003: Global analyses of sea surface temperature, sea
ice, and night marine air temperature since the late nineteenth century.
Journal of Geophysical Research: Atmospheres, 108(D14), 4407, doi:10.1029/
2002JD002670.
Reid, P.C., N.P. Holliday, and T.J. Smyth, 2001: Pulses in the eastern margin current
and warmer water off the north west European shelf linked to North Sea
ecosystem changes. Marine Ecology Progress Series, 215, 283-287.
Reid, P.C., D.G. Johns, M. Edwards, M. Starr, M. Poulin, and P. Snoeijs, 2007: A biological
consequence of reducing Arctic ice cover: arrival of the Pacific diatom
Neodenticula seminae in the North Atlantic for the first time in 800,000 years.
Global Change Biology, 13(9), 1910-1921.
Richter, I., S.K. Behera, Y. Masumoto, B. Taguchi, N. Komori, and T. Yamagata, 2010:
On the triggering of Benguela Niños: remote equatorial versus local influences.
Geophysical Research Letters, 37(20), L20604, doi:10.1029/2010GL044461.
Ridgwell, A. and D.N. Schmidt, 2010: Past constraints on the vulnerability of marine
calcifiers to massive carbon dioxide release. Nature Geoscience, 3(3), 196-
200.
Riegl, B., 2002: Effects of the 1996 and 1998 positive sea-surface temperature
anomalies on corals, coral diseases and fish in the Arabian Gulf (Dubai, UAE).
Marine Biology, 140(1), 29-40.
Riegl, B., 2003: Climate change and coral reefs: different effects in two high-latitude
areas (Arabian Gulf, South Africa). Coral Reefs, 22(4), 433-446.
Riegl, B. and W.E. Piller, 2003: Possible refugia for reefs in times of environmental
stress. International Journal of Earth Sciences, 92(4), 520-531.
Riegl, B.M., S.J. Purkis, A.S. Al-Cibahy, M.A. Abdel-Moati, and O. Hoegh-Guldberg,
2011: Present limits to heat-adaptability in corals and population-level
responses to climate extremes. PLoS ONE, 6(9), e24802, doi:10.1371/
journal.pone.0024802.
Riegl, B.M., A.W. Bruckner, G.P. Rowlands, S.J. Purkis, and P. Renaud, 2012: Red
Sea coral reef trajectories over 2 decades suggest increasing community
homogenization and decline in coral size. PLoS ONE, 7(5), e38396,
doi:10.1371/journal.pone.0038396.
Rixen, M., J.-M. Beckers, S. Levitus, J. Antonov, T. Boyer, C. Maillard, M. Fichaut, E.
Balopoulos, S. Iona, H. Dooley, M.-J. Garcia, B. Manca, A. Giorgetti, G. Manzella,
N. Mikhailov, N. Pinardi, and M. Zavatarelli, 2005: The Western Mediterranean
Deep Water: a proxy for climate change. Geophysical Research Letters, 32(12),
L12608, doi:10.1029/2005GL022702.
Robinson, J., P. Guillotreau, R. Jiménez-Toribio, F. Lantz, L. Nadzon, J. Dorizo, C. Gerry,
and F. Marsac, 2010: Impacts of climate variability on the tuna economy of
Seychelles. Climate Research, 43(3), 149-162.
Rockström, J., W. Steffen, K. Noone, A. Persson, F.S. Chapin III, E.F. Lambin, T.M.
Lenton, M. Scheffer, C. Folke, H.J. Schellnhuber, B. Nykvist, C.A. de Wit, T.
Hughes, S. van der Leeuw, H. Rodhe, S. Sorlin, P.K. Snyder, R. Costanza, U.
Svedin, M. Falkenmark, L. Karlberg, R.W. Corell, V.J. Fabry, J. Hansen, B. Walker,
D. Liverman, K. Richardson, P. Crutzen, and J.A. Foley, 2009: A safe operating
space for humanity. Nature, 461(7263), 472-475.
Rodó, X., M. Pascual, G. Fuchs, and A.S. Faruque, 2002: ENSO and cholera: a
nonstationary link related to climate change? Proceedings of the National
Academy of Sciences of the United States of America, 99(20), 12901-12906.
Rodolfo-Metalpa, R., C. Lombardi, S. Cocito, J.M. Hall-Spencer, and M.C. Gambi,
2010: Effects of ocean acidification and high temperatures on the bryozoan
Myriapora truncata at natural CO
2
vents. Marine Ecology, 31(3), 447-456.
Rodolfo-Metalpa, R., F. Houlbrèque, É. Tambutté, F. Boisson, C. Baggini, F.P. Patti, R.
Jeffree, M. Fine, A. Foggo, J.-P. Gattuso, and J.M. Hall-Spencer, 2011: Coral and
mollusc resistance to ocean acidification adversely affected by warming.
Nature Climate Change, 1(6), 308-312.
Rogelj, J., M. Meinshausen, and R. Knutti, 2012: Global warming under old and new
scenarios using IPCC climate sensitivity range estimates. Nature Climate
Change, 2, 248-253.
1728
Chapter 30 The Ocean
30
Rouault, M., B. Pohl, and P. Penven, 2010: Coastal oceanic climate change and
variability from 1982 to 2009 around South Africa. African Journal of Marine
Science, 32(2), 237-246.
Roy, C., C.D. van der Lingen, J.C. Coetzee, and J.R.E. Lutjeharms, 2007: Abrupt
environmental shift associated with changes in the distribution of Cape
anchovy Engraulis encrasicolus spawners in the southern Benguela. African
Journal of Marine Science, 29(3), 309-319.
Ruckelshaus, M., S. Doney, H. Galindo, J. Barry, F. Chan, J. Duffy, C. English, S. Gaines,
J. Grebmeier, and A. Hollowed, 2013: Securing ocean benefits for society in
the face of climate change. Marine Policy, 40, 154-159.
Rykaczewski, R.R. and J.P. Dunne, 2010: Enhanced nutrient supply to the California
Current Ecosystem with global warming and increased stratification in an earth
system model. Geophysical Research Letters, 37(21), L21606, doi:10.1029/
2010GL045019.
Sabatés, A., P. Martin, J. Lloret, and V. Raya, 2006: Sea warming and fish distribution:
the case of the small pelagic fish, Sardinella aurita, in the western
Mediterranean. Global Change Biology, 12(11), 2209-2219.
Sabine, C.L., R.A. Feely, N. Gruber, R.M. Key, K. Lee, J.L. Bullister, R. Wanninkhof,
C.S. Wong, D.W. Wallace, B. Tilbrook, F.J. Millero, T.H. Peng, A. Kozyr, T. Ono, and
A.F. Rios, 2004: The oceanic sink for anthropogenic CO
2
. Science, 305(5682),
367-371.
Sakamoto, T.T., H. Hasumi, M. Ishii, S. Emori, T. Suzuki, T. Nishimura, and A. Sumi,
2005: Responses of the Kuroshio and the Kuroshio Extension to global
warming in a high-resolution climate model. Geophysical Research Letters,
32(14), L14617, doi:10.1029/2005GL023384.
Samuelsson, M., 1996: Interannual salinity variations in the Baltic Sea during the
period 1954-1990. Continental Shelf Research, 16(11), 1463-1477.
Santana-Casiano, J.M., M. González-Dávila, and I.R. Ucha, 2009: Carbon dioxide
fluxes in the Benguela upwelling system during winter and spring: a
comparison between 2005 and 2006. Deep-Sea Research Part II: Topical
Studies in Oceanography, 56(8-10), 533-541.
Sarmiento, J.L., R. Slater, R. Barber, L. Bopp, S.C. Doney, A.C. Hirst, J. Kleypas, R. Matear,
U. Mikolajewicz, P. Monfray, V. Soldatov, S.A. Spall, and R. Stouffer, 2004:
Response of ocean ecosystems to climate warming. Global Biogeochemical
Cycles, 18(3), GB3003, doi:10.1029/2003GB002134.
Schluter, M.H., A. Merico, M. Reginatto, M. Boersma, K.H. Wiltshire, and W. Greve,
2010: Phenological shifts of three interacting zooplankton groups in relation
to climate change. Global Change Biology, 16(11), 3144-3153.
Schofield, C. and A. Arsana, 2012: Climate change and the limits of maritime
jurisdiction. In: Climate Change and the Oceans: Gauging the Legal and Policy
Currents in the Asia Pacific and Beyond [Warner, R. and C. Schofield (eds.)].
Edward Elgar Publishing, Cheltenham, UK and Northampton, MA, USA, pp.
127-152.
Schofield, O., S.M. Glenn, M.A. Moline, M. Oliver, A. Irwin, Y. Chao, and M. Arrott,
2013: Ocean observatories and information: building a global ocean observing
network. In: Earth System Monitoring, [Orcutt, J. (ed.)]. Springer, New York,
NY, USA, pp. 319-336.
Schøyen, H. and S. Bråthen, 2011: The Northern Sea Route versus the Suez
Canal: cases from bulk shipping. Journal of Transport Geography, 19(4), 977-
983.
Scott, D., C.M. Hall, and S. Gössling, 2012a: Tourism and Climate Change: Impacts,
Adaptation and Mitigation. Taylor and Francis, Abingdon, UK and New York,
NY, USA, 464 pp.
Scott, D., M.C. Simpson, and R. Sim, 2012b: The vulnerability of Caribbean coastal
tourism to scenarios of climate change related sea level rise. Journal of
Sustainable Tourism, 20(6), 883-898.
Seibel, B.A. and P.J. Walsh, 2001: Potential impacts of CO
2
injection on deep-sea
biota. Science, 294(5541), 319-320.
Serrano, E., R. Coma, M. Ribes, B. Weitzmann, M. García, and E. Ballesteros, 2013:
Rapid northward spread of a zooxanthellate coral enhanced by artificial
structures and sea warming in the western Mediterranean. PLoS ONE, 8(1),
e52739, doi:10.1371/journal.pone.0052739.
Shaffer, G., S.M. Olsen, and J.O.P. Pedersen, 2009: Long-term ocean oxygen depletion
in response to carbon dioxide emissions from fossil fuels. Nature Geoscience,
2(2), 105-109.
Sherman, K., I.M. Belkin, K.D. Friedland, J. O’Reilly, and K. Hyde, 2009: Accelerated
warming and emergent trends in fisheries biomass yields of the world’s large
marine ecosystems. AMBIO: A Journal of the Human Environment, 38(4), 215-
224.
Shoji, J., S. Toshito, K. Mizuno, Y. Kamimura, M. Hori, and K. Hirakawa, 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. International Council for the Exploration of the Sea
(ICES) Journal of Marine Science, 68(6), 1165-1169.
Signorini, S.R. and C.R. McClain, 2012: Subtropical gyre variability as seen from
satellites. Remote Sensing Letters, 3(6), 471-479.
Signorini, S.R., R.G. Murtugudde, C.R. McClain, J.R. Christian, J. Picaut, and A.J.
Busalacchi, 1999: Biological and physical signatures in the tropical and sub-
tropical Atlantic. Journal of Geophysical Research: Oceans, 104(C8), 18367-
18382.
Silverman, J., B. Lazar, and J. Erez, 2007: Effect of aragonite saturation, temperature,
and nutrients on the community calcification rate of a coral reef. Journal of
Geophysical Research: Oceans, 112(C5), C05004, doi:10.1029/2006JC003770.
Silverman, J., B. Lazar, L. Cao, K. Caldeira, and J. Erez, 2009: Coral reefs may start
dissolving when atmospheric CO
2
doubles. Geophysical Research Letters,
36(5), L05606, doi:10.1029/2008GL036282.
Simpson, S.D., S. Jennings, M.P. Johnson, J.L. Blanchard, P.J. Schon, D.W. Sims, and
M.J. Genner, 2011: Continental shelf-wide response of a fish assemblage to
rapid warming of the sea. Current Biology, 21(18), 1565-1570.
Smith, C.R., F.C. De Leo, A.F. Bernardino, A.K. Sweetman, and P.M. Arbizu, 2008:
Abyssal food limitation, ecosystem structure and climate change. Trends in
Ecology & Evolution, 23(9), 518-528.
Smith Jr., K.L. and R.S. Kaufmann, 1999: Long-term discrepancy between food supply
and demand in the deep eastern North Pacific. Science, 284(5417), 1174-
1177.
Smith Jr., K.L., H.A. Ruhl, B.J. Bett, D.S.M. Billett, R.S. Lampitt, and R.S. Kaufmann,
2009: Climate, carbon cycling, and deep-ocean ecosystems. Proceedings of
the National Academy of Sciences of the United States of America, 106(46),
19211-19218.
Smith, T.B., M.E. Brandt, J.M. Calnan, R.S. Nemeth, J. Blondeau, E. Kadison, M. Taylor,
and P. Rothenberger, 2013: Convergent mortality responses of Caribbean coral
species to seawater warming. Ecosphere, 4(7), 87, doi:10.1890/ES13-00107.1.
Southward, A., 1980: The western English Channel an inconstant ecosystem?
Nature, 285(5764), 361-366.
Southward, A.J., S.J. Hawkins, and M.T. Burrows, 1995: Seventy years’ observations of
changes in distribution and abundance of zooplankton and intertidal organisms
in the western English Channel in relation to rising sea temperature. Journal
of Thermal Biology, 20(1-2), 127-155.
Southward, A.J., O. Langmead, N.J. Hardman-Mountford, J. Aiken, G.T. Boalch, P.R.
Dando, M.J. Genner, I. Joint, M.A. Kendall, N.C. Halliday, R.P. Harris, R. Leaper,
N. Mieszkowska, R.D. Pingree, A.J. Richardson, D.W. Sims, T. Smith, A.W. Walne,
and S.J. Hawkins, 2005: Long-term oceanographic and ecological research in
the western English Channel. In: Advances in Marine Biology, Vol. 47
[Southward, A.J., P.A. Tyler, C.M. Young, and L.A. Fuiman (eds.)]. Elsevier
Academic Press, San Diego, CA, USA and London, UK, pp. 1-105.
Stenevik, E.K. and S. Sundby, 2007: Impacts of climate change on commercial fish
stocks in Norwegian waters. Marine Policy, 31(1), 19-31.
Stott, P.A., N.P. Gillett, G.C. Hegerl, D.J. Karoly, D.A. Stone, X.B. Zhang, and F. Zwiers,
2010: Detection and attribution of climate change: a regional perspective.
Wiley Interdisciplinary Reviews: Climate Change, 1(2), 192-211.
Stramma, L., G.C. Johnson, J. Sprintall, and V. Mohrholz, 2008: Expanding oxygen-
minimum zones in the tropical oceans. Science, 320(5876), 655-658.
Stramma, L., S. Schmidtko, L.A. Levin, and G.C. Johnson, 2010: Ocean oxygen minima
expansions and their biological impacts. Deep-Sea Research Part I:
Oceanographic Research Papers, 57(4), 587-595.
Stramma, L., E.D. Prince, S. Schmidtko, J. Luo, J.P. Hoolihan, M. Visbeck, D.W.R.
Wallace, P. Brandt, and A. Körtzinger, 2012: Expansion of oxygen minimum
zones may reduce available habitat for tropical pelagic fishes. Nature Climate
Change, 2(1), 33-37.
Strong, A.E., C.S. Barrientos, C. Duda, and J. Sapper, 1997: Improved satellite technique
for monitoring coral reef bleaching. In: Proceedings of the 8th International
Coral Reef Symposium 1997, Vol. 2 [Lessios, H.A. and I.G. Macintyre (eds.)].
Smithsonian Tropical Research Institute, Panama City, Panama, pp. 1495-1498.
Strong, A.E., G. Liu, W. Skirving, and C.M. Eakin, 2011: NOAA’s Coral Reef Watch
program from satellite observations. Annals of GIS, 17(2), 83-92.
Sugisaki, H., M. Nonaka, S. Ishizaki, K. Hidaka, T. Kameda, Y. Hirota, Y. Oozeki, H.
Kubota, and A. Takasuka, 2010: Status and trends of the Kuroshio region,
2003-2008. In: Marine Ecosystems of the North Pacific Ocean, 2003-2008
1729
The Ocean Chapter 30
30
[McKinnell, S.M. and M.J. Dagg (eds.)]. PICES Special Publication No. 4, North
Pacific Marine Science Organization (PICES), Sydney, BC, Canada, pp. 330-359.
Sultan, N., P. Cochonat, J.P. Foucher, and J. Mienert, 2004: Effect of gas hydrates
melting on seafloor slope instability. Marine Geology, 213(1-4), 379-401.
Sumaila, U.R., W.W.L. Cheung, V.W.Y. Lam, D. Pauly, and S. Herrick, 2011: Climate
change impacts on the biophysics and economics of world fisheries. Nature
Climate Change, 1(9), 449-456.
Sundby, S., 2000: Recruitment of Atlantic cod stocks in relation to temperature and
advection of copepod populations. Sarsia, 85(4), 277-298.
Sundby, S. and O. Nakken, 2008: Spatial shifts in spawning habitats of Arcto-
Norwegian cod related to multidecadal climate oscillations and climate
change. International Council for the Exploration of the Sea (ICES) Journal of
Marine Science, 65(6), 953-962.
Sweatman, H., S. Delean, and C. Syms, 2011: Assessing loss of coral cover on
Australia’s Great Barrier Reef over two decades, with implications for longer-
term trends. Coral Reefs, 30(2), 521-531.
Tallis, H., A. Guerry, and G.C. Daily, 2013: Ecosystem services. In: Encyclopedia of
Biodiversity, 2
nd
edn., Vol. 3 [Levin, S.A. (ed.)]. Academic Press, Waltham, UK,
pp. 96-104.
Tam, J., M.H. Taylor, V. Blaskovic, P. Espinoza, R.M. Ballón, E. Diaz, C. Wosnitza-Mendo,
J. Arguelles, S. Purca, P. Ayón, L. Quipuzcoa, D. Gutiérrez, E. Goya, N. Ochoa,
and M. Wolff, 2008: Trophic modeling of the Northern Humboldt Current
Ecosystem, Part I: comparing trophic linkages under La Niña and El Niño
conditions. Progress in Oceanography, 79(2-4), 352-365.
Tang, Q., 2009: Changing states of the Yellow Sea Large Marine Ecosystem:
anthropogenic forcing and climate impacts. In: Sustaining the World’s Large
Marine Ecosystems [Sherman, K., M.C. Aquarone, and S. Adams (eds.)].
International Union for Conservation of Nature and Natural Resources (IUCN),
Gland, Switzerland, pp. 77-88.
Tang, Q., X. Jin, J. Wang, Z. Zhuang, Y. Cui, and T. Meng, 2003: Decadal-scale variations
of ecosystem productivity and control mechanisms in the Bohai Sea. Fisheries
Oceanography, 12(4-5), 223-233.
Tang, X., F. Wang, Y. Chen, and M. Li, 2009: Warming trend in northern East China
Sea in recent four decades. Chinese Journal of Oceanology and Limnology,
27(2), 185-191.
Tåning, Å.V., 1949: On changes in the marine fauna of the North-Western Atlantic
Area, with special reference to Greenland. Rapports et Procès-Verbaux des
Réunions du Conseil Permanent International pour l’Exploration de la
Mer (CIEM) [International Council for the Exploration of the Sea (ICES)], 125,
26-29.
Tåning, Å.V., 1953: Long term changes in hydrography and fluctuations in fish stocks.
In: International Commision for the Northwest Atlantic Fisheries: Annual
Proceedings, Vol. 3, for the year 1952–53, Part 5 No. III. Issued from the
Headquarters of the Commission, Imperial Press, Ltd., Halifax, NS, Canada, pp.
69-77, icnaf.nafo.int/ap/ann-proc-v03.pdf.
Tanzil, J.T.I., B.E. Brown, A.W. Tudhope, and R.P. Dunne, 2009: Decline in skeletal
growth of the coral Porites lutea from the Andaman Sea, South Thailand
between 1984 and 2005. Coral Reefs, 28(2), 519-528.
Tasker, M.L. (ed.), 2008: The Effect of Climate Change on the Distribution and
Abundance of Marine Species in the OSPAR Maritime Area. ICES Cooperative
Research Report No. 293, International Council for the Exploration of the Sea
(ICES), Copenhagen, Denmark, 45 pp.
Taylor, G.T., F.E. Muller-Karger, R.C. Thunell, M.I. Scranton, Y. Astor, R. Varela, L.T.
Ghinaglia, L. Lorenzoni, K.A. Fanning, and S. Hameed, 2012: Ecosystem responses
in the southern Caribbean Sea to global climate change. Proceedings of the
National Academy of Sciences of the United States of America, 109(47),
19315-19320.
Taylor, M.H., J. Tam, V. Blaskovic, P. Espinoza, R.M. Ballón, C. Wosnitza-Mendo, J.
Arguelles, E. Diaz, S. Purca, N. Ochoa, P. Ayón, E. Goya, D. Gutrrez, L. Quipuzcoa,
and M. Wolff, 2008: Trophic modeling of the Northern Humboldt Current
Ecosystem, Part II: elucidating ecosystem dynamics from 1995 to 2004 with a
focus on the impact of ENSO. Progress in Oceanography, 79(2-4), 366-378.
ter Hofstede, R., J.G. Hiddink, and A.D. Rijnsdorp, 2010: Regional warming changes
fish species richness in the eastern North Atlantic Ocean. Marine Ecology
Progress Series, 414, 1-9.
Terray, L., M.E. Demory, M. Déqué, G. de Coetlogon, and E. Maisonnave, 2004:
Simulation of late-twenty-first-century changes in wintertime atmospheric
circulation over Europe due to anthropogenic causes. Journal of Climate,
17(24), 4630-4635.
Terray, L., L. Corre, S. Cravatte, T. Delcroix, G. Reverdin, and A. Ribes, 2012: Near-
surface salinity as natures rain gauge to detect human influence on the
tropical water cycle. Journal of Climate, 25(3), 958-977.
Thackeray, S.J., T.H. Sparks, M. Frederiksen, S. Burthe, P.J. Bacon, J.R. Bell, M.S. Botham,
T.M. Brereton, P.W. Bright, and L. Carvalho, 2010: Trophic level asynchrony in
rates of phenological change for marine, freshwater and terrestrial environments.
Global Change Biology, 16(12), 3304-3313.
The MerMex Group, 2011: Marine ecosystems’ responses to climatic and
anthropogenic forcings in the Mediterranean. Progress in Oceanography,
91(2), 97-166.
Thornes, J., M. Rennie, H. Marsden, and L. Chapman, 2012: Climate Change Risk
Assessment for the Transport Sector. CCRA – UK 2012 Climate Change Risk
Assessment, DEFRA Project Code GA0204, Department for the Environment,
Food and Rural Affairs (DEFRA), London, UK, 140 pp.
Thresher, R.E., B.D. Tilbrook, S. Fallon, N.C. Wilson, and J. Adkins, 2011: Effects of
chronic low carbonate saturation levels on the distribution, growth and skeletal
chemistry of deep-sea corals and other seamount megabenthos. Marine
Ecology Progress Series, 442, 87-99.
Tian, Y.J., H. Kidokoro, T. Watanabe, Y. Igeta, H. Sakaji, and S. Ino, 2012: Response of
yellowtail, Seriola quinqueradiata, a key large predatory fish in the Japan Sea,
to sea water temperature over the last century and potential effects of global
warming. Journal of Marine Systems, 91(1), 1-10.
Tokinaga, H. and S.-P. Xie, 2011: Weakening of the equatorial Atlantic cold tongue
over the past six decades. Nature Geoscience, 4(4), 222-226.
Toresen, R. and O.J. Østvedt, 2000: Variation in abundance of Norwegian spring-
spawning herring (Clupea harengus, Clupeidae) throughout the 20
t
h
century
and the influence of climatic fluctuations. Fish and Fisheries, 1(3), 231-256.
Touratier, F. and C. Goyet, 2011: Impact of the Eastern Mediterranean Transient on
the distribution of anthropogenic CO
2
and first estimate of acidification for
the Mediterranean Sea. Deep-Sea Research Part I: Oceanographic Research
Papers, 58(1), 1-15.
Trenberth, K.E., 1990: Recent observed interdecadal climate changes in the Northern
Hemisphere. Bulletin of the American Meteorological Society, 71(7), 988-993.
Trenberth, K.E. and D.J. Shea, 2006: Atlantic hurricanes and natural variability in 2005.
Geophysical Research Letters, 33(12), L12704, doi:10.1029/2006GL026894.
Tsamenyi, M. and Q. Hanich, 2012: Fisheries jurisdiction under the Law of the Sea
Convention: rights and obligations in maritime zones under the sovereignty
of Coastal States. The International Journal of Marine and Coastal Law, 27(4),
783-793.
Tsikliras, A., 2008: Climate-related geographic shift and sudden population increase
of a small pelagic fish (Sardinella aurita) in the eastern Mediterranean Sea.
Marine Biology Research, 4(6), 477-481.
Turley, C. and K. Boot, 2011: The ocean acidification challenges facing science and
society. In: Ocean Acidification [Gattuso, J.-P. and L. Hansson (eds.)]. Oxford
University Press, Oxford, UK, pp. 249-271.
Turner, R.E. and N.N. Rabalais, 1994: Coastal eutrophication near the Mississippi
river delta. Nature, 368(6472), 619-621.
Turner, R.E., N.N. Rabalais, and D. Justic, 2008: Gulf of Mexico hypoxia: alternate
states and a legacy. Environmental Science & Technology, 42(7), 2323-2327.
Turton, S., T. Dickson, W. Hadwen, B. Jorgensen, T. Pham, D. Simmons, P. Tremblay,
and R. Wilson, 2010: Developing an approach for tourism climate change
assessment: evidence from four contrasting Australian case studies. Journal
of Sustainable Tourism, 18(3), 429-447.
Tyler, W.B., K.T. Briggs, D.B. Lewis, and R.G. Ford, 1993: Seabird distribution and
abundance in relation to oceanographic processes in the California Current
System. In: The Status, Ecology and Conservation of Marine Birds of the North
Pacific [Vermeer, K., K.T. Briggs, K.H. Morgan, and D. Siegel-Causey (eds.)].
Special Publication, Canadian Wildlife Series, Canadian Wildlife Service,
Environment Canada, Ottawa, ON, Canada, pp. 48-60.
Ukrainskii, V. and Y.I. Popov, 2009: Climatic and hydrophysical conditions of the
development of hypoxia in waters of the northwest shelf of the Black Sea.
Physical Oceanography, 19(3), 140-150.
UNCTAD, 2009a: Review of Maritime Transport 2009. Report by the United Nations
Conference on Trade and Development (UNCTAD) Secretariat, United Nations,
New York, NY, USA and Geneva, Switzerland, 221 pp.
UNCTAD, 2009b: Multi-Year Expert Meeting on Transport and Trade Facilitation:
Maritime Transport and the Climate Change Challenge, 16-18 February 2009,
Geneva. Summary of Proceedings. United Nations Conference on Trade and
Development (UNCTAD), United Nations, Geneva, Switzerland, 52 pp.
1730
Chapter 30 The Ocean
30
UNWTO and UNEP, 2008: Climate Change and Tourism: Responding to Global
Challenges. World Tourism Organization (UNWTO), Madrid, Spain and the
United Nations Environment Programme (UNEP), Paris, France, 256 pp.
Van Dover, C.L., 2012: Hydrothermal vent ecosystems and conservation. Oceanography,
25(1), 313-316.
van Hooidonk, R., J.A. Maynard, and S. Planes, 2013: Temporary refugia for coral
reefs in a warming world. Nature Climate Change, 3, 508-511.
van Vuuren, D.P., J. Edmonds, M. Kainuma, K. Riahi, A. Thomson, K. Hibbard, G.C.
Hurtt, T. Kram, V. Krey, J.F. Lamarque, T. Masui, M. Meinshausen, N. Nakicenovik,
S.J. Smith, and S.K. Rose, 2011: The representative concentration pathways:
an overview. Climatic Change, 109(1-2), 5-31.
Vantrepotte, V. and F. lin, 2011: Inter-annual variations in the SeaWiFS global
chlorophyll a concentration (1997-2007). Deep-Sea Research Part I:
Oceanographic Research Papers, 58(4), 429-441.
Vargas-Ángel, B., E.E. Looney, O.J. Vetter, and E.F. Coccagna, 2011: Severe, widespread
El Niño-associated coral bleaching in the US Phoenix Islands. Bulletin of
Marine Science, 87(3), 623-638.
Vargas-Yáñez, M., F. Moya, M.C. Garcia-Martinez, E. Tel, P. Zunino, F. Plaza, J. Salat,
J. Pascual, J.L. López-Jurado, and M. Serra, 2010: Climate change in the
Western Mediterranean Sea 1900-2008. Journal of Marine Systems, 82(3),
171-176.
Vecchi, G.A., B.J. Soden, A.T. Wittenerg, I.M. Held, A. Leetmaa, and M.J. Harrison, 2006:
Weakening of tropical Pacific atmospheric circulation due to anthropogenic
forcing. Nature, 441(7089), 73-76.
Veitch, J., 2007: The Changing State of the Benguela Current Large Marine Ecosystem:
Expert Workshop on Climate Change and Variability and Impacts Thereof in the
BCLME Region, 15-16 May 2007. Workshop Report, Global Environment Facility
(GEF), United Nations Development Program (UNDP), Washington, DC, USA,
55 pp.
Veron, J.E.N., L.M. Devantier, E. Turak, A.L. Green, S. Kininmonth, M. Stafford-Smith,
and N. Peterson, 2009: Delineating the coral triangle. Galaxea, Journal of Coral
Reef Studies, 11(2), 91-100.
Wang, C. and S. Dong, 2010: Is the basin-wide warming in the North Atlantic Ocean
related to atmospheric carbon dioxide and global warming? Geophysical
Research Letters, 37(8), L08707, doi:10.1029/2010GL042743.
Wang, M., J.E. Overland, D.B. Percival, and H.O. Mofjeld, 2006: Change in the arctic
influence on Bering Sea climate during the twentieth century. International
Journal of Climatology, 26(4), 531-539.
Wang, M., N.A. Bond, and J.E. Overland, 2007a: Comparison of atmospheric forcing in
four sub-arctic seas. Deep-Sea Research Part II: Topical Studies in Oceanography,
54(23-26), 2543-2559.
Wang, M., J.E. Overland, V. Kattsov, J.E. Walsh, X. Zhang, and T. Pavlova, 2007b:
Intrinsic versus forced variation in coupled climate model simulations
over the Arctic during the twentieth century. Journal of Climate, 20(6), 1093-
1107.
Wang, X., J.R. Christian, R. Murtugudde, and A.J. Busalacchi, 2006: Spatial and
temporal variability in new production in the equatorial Pacific during 1980-
2003: physical and biogeochemical controls. Deep-Sea Research Part II: Topical
Studies in Oceanography, 53(5-7), 677-697.
Warner, R., 2012: Oceans beyond boundaries: environmental assessment frameworks.
International Journal of Marine and Coastal Law, 27(2), 481-499.
Warner, R. and C. Schofield (eds.), 2012: Climate Change and the Oceans: Gauging
the Legal and Policy Currents in the Asia Pacific and Beyond. Edward Elgar
Publishing, Cheltenham, UK and Northampton, MA, USA, 296 pp.
Waycott, M., C. Collier, K. McMahon, P. Ralph, L. McKenzie, J. Udy, and A. Grech,
2007: Vulnerability of seagrasses in the Great Barrier Reef to climate change.
In: Climate Change and the Great Barrier Reef: A Vulnerability Assessment
[Johnson, J.E. and P.A. Marshall (eds.)]. Great Barrier Reef Marine Park
Authority (GBRMPA), Townsville, QLD, Australia and Department of the
Environment and Heritage, Australian Greenhouse Office, Canberra, ACT,
Australia, pp. 193-236.
Waycott, M., L.J. McKenzie, J.E. Mellors, J.E. Ellison, M.T. Sheaves, C. Collier, A.-M.
Schwarz, A. Webb, J.E. Johnson, and C.E. Payri, 2011: Vulnerability of mangroves,
seagrasses and intertidal flats in the tropical Pacific to climate change. In:
Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change
[Bell, J.D., J.E. Johnson, and A.J. Hobday (eds.)]. Secretariat of the Pacific
Community, Noumea, New Caledonia, pp. 297-368.
Weeks, S.J., B. Currie, and A. Bakun, 2002: Satellite imaging: massive emissions of
toxic gas in the Atlantic. Nature, 415(6871), 493-494.
West, J.M., S.H. Julius, P. Kareiva, C. Enquist, J.J. Lawler, B. Petersen, A.E. Johnson, and
M.R. Shaw, 2009: US natural resources and climate change: concepts and
approaches for management adaptation. Environmental Management, 44(6),
1001-1021.
Westmacott, S., H. Cesar, and L. Pet-Soede, 2000: Socio-economic assessment of the
impacts of the 1998 coral reef bleaching in the Indian Ocean: a summary. In:
Coral Reef Degradation in the Indian Ocean: Status Report 2000 [Souter, D.,
D. Obura, and O. Lindén (eds.)]. Coral Reef Degradation in the Indian Ocean
Programme (CORDIO), SAREC Marine Science Program, Department of
Zoology, Stockholm University, Stockholm, Sweden, pp. 143-159.
Wethey, D.S. and S.A. Woodin, 2008: Ecological hindcasting of biogeographic
responses to climate change in the European intertidal zone. Hydrobiologia,
606(1), 139-151.
Whitney, F.A., H.J. Freeland, and M. Robert, 2007: Persistently declining oxygen levels
in the interior waters of the eastern subarctic Pacific. Progress in Oceanography,
75(2), 179-199.
Wild, M., 2009: Global dimming and brightening: a review. Journal of Geophysical
Research: Atmospheres, 114(D10), D00D16, doi:10.1029/2008JD011470.
Wilderbuer, T.K., A.B. Hollowed, W.J. Ingraham, P.D. Spencer, M.E. Conners, N.A. Bond,
and G.E. Walters, 2002: Flatfish recruitment response to decadal climatic
variability and ocean conditions in the eastern Bering Sea. Progress in
Oceanography, 55(1-2), 235-247.
Wilkinson, C., 2004. Status of Coral Reefs of the World: 2004. Australian Institute of
Marine Science, Townsville, QLD, Australia, 264 pp.
Wilkinson, C. and G. Hodgson, 1999: Coral reefs and the 1997-1998 mass bleaching
and mortality. Nature & Resources, 35(2), 16-25.
Wiltshire, K.H. and B.F.J. Manly, 2004: The warming trend at Helgoland Roads, North
Sea: phytoplankton response. Helgoland Marine Research, 58(4), 269-273.
Wiltshire, K.H., A.M. Malzahn, K. Wirtz, W. Greve, S. Janisch, P. Mangelsdorf, B.F.J.
Manly, and M. Boersma, 2008: Resilience of North Sea phytoplankton spring
bloom dynamics: an analysis of long-term data at Helgoland Roads. Limnology
and Oceanography, 53(4), 1294-1302.
Wong, A.P.S., N.L. Bindoff, and J.A. Church, 2001: Freshwater and heat changes in
the North and South Pacific Oceans between the 1960s and 1985-94. Journal
of Climate, 14(7), 1613-1633.
Wong, G.T.F., S.Y. Chao, Y.H. Li, and F.K. Shiah, 2000: The Kuroshio edge exchange
processes (KEEP) study an introduction to hypotheses and highlights.
Continental Shelf Research, 20(4-5), 335-347.
Wood, K.R. and J.E. Overland, 2010: Early 20
th
century Arctic warming in retrospect.
International Journal of Climatology, 30(9), 1269-1279.
Woodworth-Jefcoats, P.A., J.J. Polovina, J.P. Dunne, and J.L. Blanchard, 2013:
Ecosystem size structure response to 21
st
century climate projection: large fish
abundance decreases in the central North Pacific and increases in the
California Current. Global Change Biology, 19(3), 724-733.
World Bank and FAO, 2008: The Sunken Billions: The Economic Justification for
Fisheries Reform. Architecture and Rural Development Series, The International
Bank for Reconstruction and Development / The World Bank and the Food and
Agriculture Organization of the United Nations (FAO), World Bank Publications,
Washington, DC, USA, 100 pp.
Worm, B., R. Hilborn, J.K. Baum, T.A. Branch, J.S. Collie, C. Costello, M.J. Fogarty, E.A.
Fulton, J.A. Hutchings, S. Jennings, O.P. Jensen, H.K. Lotze, P.M. Mace, T.R.
McClanahan, C. Minto, S.R. Palumbi, A.M. Parma, D. Ricard, A.A. Rosenberg,
R. Watson, and D. Zeller, 2009: Rebuilding global fisheries. Science, 325(5940),
578-585.
WTTC, 2012: The Economic Impact of Travel & Tourism 2012. World Travel & Tourism
Council (WTTC), London, UK, 20 pp.
Wu, L., W. Cai, L. Zhang, H. Nakamura, A. Timmermann, T. Joyce, M.J. McPhaden, M.
Alexander, B. Qiu, M. Visbecks, P. Chang, and B. Giese, 2012: Enhanced
warming over the global subtropical western boundary currents. Nature
Climate Change, 2(3), 161-166.
Wunsch, C. and P. Heimbach, 2008: How long to oceanic tracer and proxy equilibrium?
Quaternary Science Reviews, 27(7-8), 637-651.
Yatsu, A., K.Y. Aydin, J.R. King, G.A. McFarlane, S. Chiba, K. Tadokoro, M. Kaeriyama,
and Y. Watanabe, 2008: Elucidating dynamic responses of North Pacific fish
populations to climatic forcing: influence of life-history strategy. Progress in
Oceanography, 77(2-3), 252-268.
Yatsu, A., S. Chiba, Y. Yamanaka, S. Ito, Y. Shimizu, M. Kaeriyama, and Y. Watanabe, 2013:
Climate forcing and the Kuroshio/Oyashio ecosystem. International Council for
the Exploration of the Sea (ICES) Journal of Marine Science, 70(5), 922-933.
1731
The Ocean Chapter 30
30
Ye, S. and X. Huang, 2003: HABs in East China Sea: surveillance and monitoring.
Marine Environmental Science/Haiyang Huanjing Kexue, 22(2), 10-14.
Yoder, J.A., C.R. Mcclain, G.C. Feldman, and W.E. Esaias, 1993: Annual cycles of
phytoplankton chlorophyll concentrations in the global ocean a satellite
view. Global Biogeochemical Cycles, 7(1), 181-193.
Yu, J.Y. and H.Y. Kao, 2007: Decadal changes of ENSO persistence barrier in SST and
ocean heat content indices: 1958-2001. Journal of Geophysical Research:
Atmospheres, 112(D13), D13106, doi:10.1029/2006JD007654.
Yu, K.-F., J.-X. Zhao, Q. Shi, T.-G. Chen, P.-X. Wang, K.D. Collerson, and T.-S. Liu, 2006:
U-series dating of dead Porites corals in the South China Sea: evidence for
episodic coral mortality over the past two centuries. Quaternary Geochronology,
1(2), 129-141.
Yunev, O.A., J. Carstensen, S. Moncheva, A. Khaliulin, G. Ærtebjerg, and S. Nixon,
2007: Nutrient and phytoplankton trends on the western Black Sea shelf in
response to cultural eutrophication and climate changes. Estuarine Coastal
and Shelf Science, 74(1-2), 63-76.
Zeebe, R.E. and A. Ridgwell, 2011: Past changes in ocean carbonate chemistry. In:
Ocean Acidification [Gattuso, J.-P. and L. Hansson (eds.)]. Oxford University
Press, Oxford, UK, pp. 21-66.
Zeeberg, J., A. Corten, P. Tjoe-Awie, J. Coca, and B. Hamady, 2008: Climate modulates
the effects of Sardinella aurita fisheries off Northwest Africa. Fisheries
Research, 89(1), 65-75.
Zervas, C.E., 2009. Sea Level Variations of the United States, 1854-2006. Technical
Report NOS CO-OPS 053, National Oceanic and Atmospheric Administration
(NOAA), Silver Spring, MD, USA, 194 pp.
Zhang, Q., Y. Hou, and T. Yan, 2011: Inter-annual and inter-decadal variability of
Kuroshio heat transport in the East China Sea. International Journal of
Climatology, 32(4), 481-488.
Zhang, X., J.A. Church, S.M. Platten, and D. Monselesan, 2014: Projection of subtropical
gyre circulation and associated sea level changes in the Pacific based on
CMIP3 climate models. Climate Dynamics, 43, 131-144, doi:10.1007/s00382-
013-1902-x.
Zhang, X., S. Shang, C. Zhang, Z. Lu, S. Shang, and X. Li, 2005: Potential influence of
sea surface temperature on the interannual fluctuation of the catch and the
distribution of chub mackerel and round scad in the Minnan-Taiwan Bank
Fishing Ground, China. Chinese Marine Science Bulletin, 24(4), 91-96.
Zhao, M., K. Yu, Q. Zhang, Q. Shi, and G.J. Price, 2012: Long-term decline of a fringing
coral reef in the Northern South China Sea. Journal of Coastal Research, 28(5),
1088-1099.