1371
25
Australasia
Coordinating Lead Authors:
Andy Reisinger (New Zealand), Roger L. Kitching (Australia)
Lead Authors:
Francis Chiew (Australia), Lesley Hughes (Australia), Paul C.D. Newton (New Zealand),
Sandra S. Schuster (Australia), Andrew Tait (New Zealand), Penny Whetton (Australia)
Contributing Authors:
Jon Barnett (Australia), Susanne Becken (New Zealand), Paula Blackett (New Zealand),
Sarah Boulter (Australia), Andrew Campbell (Australia), Daniel Collins (New Zealand),
Jocelyn Davies (Australia), Keith Dear (Australia), Stephen Dovers (Australia), Kyla Finlay
(Australia), Bruce Glavovic (New Zealand), Donna Green (Australia), Don Gunasekera
(Australia), Simon Hales (New Zealand), John Handmer (Australia), Garth Harmsworth
(New Zealand), Alistair Hobday (Australia), Mark Howden (Australia), Graeme Hugo
(Australia), Sue Jackson (Australia), David Jones (Australia), Darren King (New Zealand),
Miko Kirschbaum (New Zealand), Jo Luck (Australia), Yiheyis Maru (Australia), Jan McDonald
(Australia), Kathy McInnes (Australia), Johanna Mustelin (Australia), Barbara Norman
(Australia), Grant Pearce (New Zealand), Susan Peoples (New Zealand), Ben Preston (USA),
Joseph Reser (Australia), Penny Reyenga (Australia), Mark Stafford-Smith (Australia),
Xiaoming Wang (Australia), Leanne Webb (Australia)
Review Editors:
Blair Fitzharris (New Zealand), David Karoly (Australia)
This chapter should be cited as:
Reisinger
, A., R.L. Kitching, F. Chiew, L. Hughes, P.C.D. Newton, S.S. Schuster, A. Tait, and P. Whetton, 2014:
Australasia. 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. 1371-1438.
25
1372
Executive Summary ......................................................................................................................................................... 1374
25.1. Introduction and Major Conclusions from Previous Assessments ........................................................................ 1377
25.2. Observed and Projected Climate Change ............................................................................................................. 1377
25.3. Socioeconomic Trends Influencing Vulnerability and Adaptive Capacity .............................................................. 1379
25.3.1. Economic, Demographic, and Social Trends .................................................................................................................................... 1379
25.3.2. Use and Relevance of Socioeconomic Scenarios in Adaptive Capacity/Vulnerability Assessments .................................................. 1382
25.4. Cross-Sectoral Adaptation: Approaches, Effectiveness, and Constraints .............................................................. 1382
25.4.1. Frameworks, Governance, and Institutional Arrangements ............................................................................................................. 1382
25.4.2. Constraints on Adaptation and Leading Practice Models ............................................................................................................... 1382
Box 25-1. Coastal Adaptation—Planning and Legal Dimensions .............................................................................................. 1384
25.4.3. Sociocultural Factors Influencing Impacts of and Adaptation to Climate Change ........................................................................... 1385
25.5. Freshwater Resources ........................................................................................................................................... 1387
25.5.1. Observed Impacts ........................................................................................................................................................................... 1387
25.5.2. Projected Impacts ........................................................................................................................................................................... 1387
25.5.3. Adaptation ..................................................................................................................................................................................... 1389
Box 25-2. Adaptation through Water Resources Policy and Management in Australia ............................................................ 1389
25.6. Natural Ecosystems ............................................................................................................................................... 1390
25.6.1. Inland Freshwater and Terrestrial Ecosystems ................................................................................................................................. 1390
25.6.1.1. Observed Impacts ............................................................................................................................................................ 1390
25.6.1.2. Projected Impacts ............................................................................................................................................................ 1390
25.6.1.3. Adaptation ...................................................................................................................................................................... 1391
25.6.2. Coastal and Ocean Ecosystems ...................................................................................................................................................... 1392
25.6.2.1. Observed Impacts ............................................................................................................................................................ 1392
25.6.2.2. Projected Impacts ............................................................................................................................................................ 1392
25.6.2.3. Adaptation ...................................................................................................................................................................... 1392
Box 25-3. Impacts of a Changing Climate in Natural and Managed Ecosystems .......................................................... 1394
25.7. Major Industries .................................................................................................................................................... 1393
25.7.1. Production Forestry ......................................................................................................................................................................... 1393
25.7.1.1. Observed and Projected Impacts ..................................................................................................................................... 1395
25.7.1.2. Adaptation ...................................................................................................................................................................... 1396
25.7.2. Agriculture ...................................................................................................................................................................................... 1396
25.7.2.1. Projected Impacts and Adaptation—Livestock Systems .................................................................................................. 1396
25.7.2.2. Projected Impacts and Adaptation—Cropping ................................................................................................................ 1397
Table of Contents
1373
Australasia Chapter 25
25
Box 25-4. Biosecurity ..................................................................................................................................................... 1397
Box 25-5. Climate Change Vulnerability and Adaptation in Rural Areas ....................................................................... 1398
25.7.2.3. Integrated Adaptation Perspectives ................................................................................................................................. 1399
Box 25-6. Climate Change and Fire ................................................................................................................................ 1400
25.7.3. Mining ............................................................................................................................................................................................ 1399
25.7.4. Energy Supply, Transmission, and Demand ..................................................................................................................................... 1400
25.7.5. Tourism ........................................................................................................................................................................................... 1401
25.7.5.1. Projected Impacts ............................................................................................................................................................ 1401
25.7.5.2. Adaptation ...................................................................................................................................................................... 1401
25.8. Human Society ...................................................................................................................................................... 1402
25.8.1. Human Health ................................................................................................................................................................................ 1402
25.8.1.1. Observed Impacts ............................................................................................................................................................ 1402
25.8.1.2. Projected Impacts ............................................................................................................................................................ 1402
Box 25-7. Insurance as Climate Risk Management Tool ................................................................................................ 1403
Box 25-8. Changes in Flood Risk and Management Responses ..................................................................................... 1404
25.8.1.3. Adaptation ...................................................................................................................................................................... 1405
25.8.2. Indigenous Peoples ......................................................................................................................................................................... 1405
25.8.2.1. Aboriginal and Torres Strait Islanders .............................................................................................................................. 1405
25.8.2.2. New Zealand Māori ......................................................................................................................................................... 1405
25.9. Interactions among Impacts, Adaptation, and Mitigation Responses .................................................................. 1406
Box 25-9. Opportunities, Constraints, and Challenges to Adaptation in Urban Areas .............................................................. 1406
25.9.1. Interactions among Local-Level Impacts, Adaptation, and Mitigation Responses ........................................................................... 1406
25.9.2. Intra- and Inter-regional Flow-on Effects among Impacts, Adaptation, and Mitigation .................................................................. 1408
Box 25-10. Land-based Interactions among Climate, Energy, Water, and Biodiversity ............................................................. 1409
25.10.Synthesis and Regional Key Risks ......................................................................................................................... 1410
25.10.1. Economy-wide Impacts and Potential of Mitigation to Reduce Risks ............................................................................................ 1410
25.10.2. Regional Key Risks as a Function of Mitigation and Adaptation ................................................................................................... 1410
25.10.3. The Role of Adaptation in Managing Key Risks, and Adaptation Limits ......................................................................................... 1412
25.11.Filling Knowledge Gaps to Improve Management of Climate Risks .................................................................... 1412
References ....................................................................................................................................................................... 1414
Frequently Asked Questions
25.1: How can we adapt to climate change if projected future changes remain uncertain? ................................................................... 1386
25.2: What are the key risks from climate change to Australia and New Zealand? ................................................................................. 1412
1374
Chapter 25 Australasia
25
Executive Summary
The regional climate is changing (very high confidence). The region continues to demonstrate long-term trends toward higher surface air
and sea surface temperatures, more hot extremes and fewer cold extremes, and changed rainfall patterns. Over the past 50 years, increasing
greenhouse gas concentrations have contributed to rising average temperature in Australia (high confidence) and New Zealand (medium
confidence) and decreasing rainfall in southwestern Australia (high confidence). {25.2; Table 25-1}
Warming is projected to continue through the 21st century (virtually certain) along with other changes in climate. Warming is
expected to be associated with rising snow lines (very high confidence), more frequent hot extremes, less frequent cold extremes (high
confidence), and increasing extreme rainfall related to flood risk in many locations (medium confidence). Annual average rainfall is expected to
decrease in southwestern Australia (high confidence) and elsewhere in most of far southern Australia and the northeast South Island and
northern and eastern North Island of New Zealand (medium confidence), and to increase in other parts of New Zealand (medium confidence).
Tropical cyclones are projected to increase in intensity but remain similar or decrease in numbers (low confidence), and fire weather is projected
to increase in most of southern Australia (high confidence) and many parts of New Zealand (medium confidence). Regional sea level rise will
very likely exceed the historical rate (1971–2010), consistent with global mean trends. {25.2; Table 25-1; Box 25-6; WGI AR5 13.5-6}
Uncertainty in projected rainfall changes remains large for many parts of Australia and New Zealand, which creates significant
challenges for adaptation. For example, projections for average annual runoff in far southeastern Australia range from little change to a
40% decline for 2°C global warming above current levels. The dry end of these scenarios would have severe implications for agriculture, rural
livelihoods, ecosystems, and urban water supply, and would increase the need for transformational adaptation (high confidence). {25.2, 25.5.1,
25.6.1, 25.7.2; Boxes 25-2, 25-5}
Recent extreme climatic events show significant vulnerability of some ecosystems and many human systems to current climate
variability (very high confidence), and the frequency and/or intensity of such events is projected to increase in many locations
(medium to high confidence).
For example, high sea surface temperatures have repeatedly bleached coral reefs in northeastern Australia
(since the late 1970s) and more recently in western Australia. Recent floods in Australia and New Zealand caused severe damage to infrastructure
and settlements and 35 deaths in Queensland alone (2011); the Victorian heat wave (2009) increased heat-related morbidity and was associated
with more than 300 excess deaths, while intense bushfires destroyed more than 2000 buildings and led to 173 deaths; and widespread drought
in southeast Australia (1997–2009) and many parts of New Zealand (2007–2009; 2012–2013) resulted in substantial economic losses (e.g.,
regional gross domestic product (GDP) in the southern Murray-Darling Basin was below forecast by about 5.7% in 2007–2008, and New
Zealand lost about NZ$3.6 billion in direct and off-farm output in 2007–2009). {25.6.2, 25.8.1; Table 25-1; Boxes 25-5, 25-6, 25-8}
Without adaptation, further changes in climate, atmospheric carbon dioxide (CO
2
), and ocean acidity are projected to have
substantial impacts on water resources, coastal ecosystems, infrastructure, health, agriculture, and biodiversity (high confidence).
Freshwater resources are projected to decline in far southwest and far southeast mainland Australia (high confidence) and for rivers originating
in the northeast of the South Island and east and north of the North Island of New Zealand (medium confidence). Rising sea levels and increasing
heavy rainfall are projected to increase erosion and inundation, with consequent damages to many low-lying ecosystems, infrastructure, and
housing; increasing heat waves will increase risks to human health; rainfall changes and rising temperatures will shift agricultural production
zones; and many native species will suffer from range contractions and some may face local or even global extinction. {25.5.1, 25.6.1-2, 25.7.2,
25.7.4; Boxes 25-1, 25-5, 25-8}
Some sectors in some locations have the potential to benefit from projected changes in climate and increasing atmospheric CO
2
(high confidence). Examples include reduced winter mortality (low confidence), reduced energy demand for winter heating in New Zealand
and southern parts of Australia, and forest growth in cooler regions except where soil nutrients or rainfall are limiting. Spring pasture growth in
cooler regions would also increase and be beneficial for animal production if it can be utilized. {25.7.1-2, 25.7.4, 25.8.1}
Adaptation is already occurring and adaptation planning is becoming embedded in some planning processes, albeit mostly at the
conceptual rather than implementation level (high confidence).
Many solutions for reducing energy and water consumption in urban
areas with co-benefits for climate change adaptation (e.g., greening cities and recycling water) are already being implemented. Planning for
1375
25
Australasia Chapter 25
reduced water availability in southern Australia and for sea level rise in both countries is becoming adopted widely, although implementation
of specific policies remains piecemeal, subject to political changes, and open to legal challenges. {25.4; Boxes 25-1, 25-2, 25-9}
Adaptive capacity is generally high in many human systems, but implementation faces major constraints, especially for
transformational responses at local and community levels (high confidence). Efforts to understand and enhance adaptive capacity and
adaptation processes have increased since the AR4, particularly in Australia. Constraints on implementation arise from: absence of a consistent
information base and uncertainty about projected impacts; limited financial and human resources to assess local risks and to develop and
implement effective policies and rules; limited integration of different levels of governance; lack of binding guidance on principles and priorities;
different attitudes toward the risks associated with climate change; and different values placed on objects and places at risk. {25.4, 25.10.3;
Table 25-2; Box 25-1}
Indigenous peoples in both Australia and New Zealand have higher than average exposure to climate change because of a heavy
reliance on climate-sensitive primary industries and strong social connections to the natural environment, and face particular
constraints to adaptation (medium confidence).
Social status and representation, health, infrastructure and economic issues, and engage-
ment with natural resource industries constrain adaptation and are only partly offset by intrinsic adaptive capacity (high confidence). Some
proposed responses to climate change may provide economic opportunities, particularly in New Zealand related to forestry. Torres Strait
communities are vulnerable even to small sea level rises (high confidence). {25.3, 25.8.2}
We identify eight regional key risks during the 21st century based on the severity of potential impacts for different levels of
warming, uniqueness of the systems affected, and adaptation options (high confidence). These risks differ in the degree to which they
can be managed via adaptation and mitigation, and some are more likely to be realized than others, but all warrant attention from a risk-
management perspective.
Some potential impacts can be delayed but now appear very difficult to avoid entirely, even with globally effective mitigation and planned
adaptation:
Significant change in community composition and structure of coral reef systems in Australia, driven by increasing sea surface
temperatures and ocean acidification; the ability of corals to adapt naturally to rising temperatures and acidification appears limited
and insufficient to offset the detrimental effects. {25.6.2, 30.5; Box CC-CR}
Loss of montane ecosystems and some native species in Australia, driven by rising temperatures and snow lines, increased fire risk, and
drying trends; fragmentation of landscapes, limited dispersal, and limited rate of evolutionary change constrain adaptation options. {25.6.1}
Some impacts have the potential to be severe but can be reduced substantially by globally effective mitigation combined with adaptation,
with the need for transformational adaptation increasing with the rate and magnitude of climate change:
Increased frequency and intensity of flood damage to settlements and infrastructure in Australia and New Zealand, driven by increasing
extreme rainfall although the amount of change remains uncertain; in many locations, continued reliance on increased protection
alone would become progressively less feasible. {25.4.2, 25.10.3; Table 25-1; Box 25-8}
Constraints on water resources in southern Australia, driven by rising temperatures and reduced cool-season rainfall; integrated
responses encompassing management of supply, recycling, water conservation, and increased efficiency across all sectors are available
and some are being implemented in areas already facing shortages. {25.2, 25.5.2; Box 25-2}
Increased morbidity, mortality, and infrastructure damages during heat waves in Australia, resulting from increased frequency and
magnitude of extreme high temperatures; vulnerable populations include the elderly and those with existing chronic diseases; population
increases and aging trends constrain effectiveness of adaptation responses. {25.8.1}
Increased damages to ecosystems and settlements, economic losses, and risks to human life from wildfires in most of southern
Australia and many parts of New Zealand
, driven by rising temperatures and drying trends; local planning mechanisms, building
design, early warning systems, and public education can assist with adaptation and are being implemented in regions that have
experienced major events. {25.2, 25.6.1, 25.7.1; Table 25-1; Box 25-6}
For some impacts, severity depends on changes in climate variables that span a particularly large range, even for a given global temperature
change. The most severe changes would present major challenges if realized:
Increasing risks to coastal infrastructure and low-lying ecosystems in Australia and New Zealand from continuing sea level rise, with
widespread damages toward the upper end of projected sea level rise ranges
; managed retreat is a long-term adaptation strategy for
1376
Chapter 25 Australasia
25
human systems but options for some natural ecosystems are limited owing to the rapidity of change and lack of suitable space for
landward migration. Risks from sea level rise continue to increase beyond 2100 even if temperatures are stabilized. {25.4.2, 25.6.1-2;
Table 25-1; Box 25-1; WGI AR5 13.5}
Significant reduction in agricultural production in the Murray-Darling Basin and far southeastern and southwestern Australia if scenarios
of severe drying are realized; more efficient water use, allocation, and trading would increase the resilience of systems in the near term
but cannot prevent significant reductions in agricultural production and severe consequences for ecosystems and some rural communities
at the dry end of the projected changes. {25.2, 25.5.2, 25.7.2; Boxes 25-2, 25-5}
Significant synergies and trade-offs exist between alternative adaptation responses, and between mitigation and adaptation
responses; interactions occur both within Australasia and between Australasia and the rest of the world (very high confidence).
Increasing efforts to mitigate and adapt to climate change imply an increasing complexity of interactions, particularly at the intersections
among water, energy, and biodiversity, but tools to understand and manage these interactions remain limited. Flow-on effects from climate
change impacts and responses outside Australasia have the potential to outweigh some of the direct impacts within the region, particularly
economic impacts on trade-intensive sectors such as agriculture (medium confidence) and tourism (limited evidence, high agreement), but they
remain among the least explored issues. {25.7.5, 25.9.1-2; Box 25-10}
Understanding of future vulnerability of human and mixed human-natural systems to climate change remains limited due to
incomplete consideration of socioeconomic dimensions (very high confidence).
Future vulnerability will depend on factors such as
wealth and its distribution across society, patterns of aging, access to technology and information, labor force participation, societal values,
and mechanisms and institutions to resolve conflicts. These dimensions have received only limited attention and are rarely included in
vulnerability assessments, and frameworks to integrate social, psychological, and cultural dimensions of vulnerability with biophysical impacts
and economic losses are lacking. In addition, conclusions for New Zealand in many sectors, even for biophysical impacts, are based on limited
studies that often use a narrow set of assumptions, models, and data and hence have not explored the full range of potential outcomes.
{25.3-4, 25.11}
1377
Australasia Chapter 25
25
25.1. Introduction and Major Conclusions
from Previous Assessments
Australasia is defined here as lands, territories, offshore waters, and
o
ceanic islands of the exclusive economic zones of Australia and New
Zealand. Both countries are relatively wealthy, with export-led economies.
Both have Westminster-style political systems and have a relatively
recent history of non-indigenous settlement (Australia in the late 18th,
New Zealand in the early 19th century). Both retain significant indigenous
populations.
Principal findings from the IPCC Fourth Assessment Report (AR4) for
the region were (Hennessy et al., 2007):
Consistent with global trends, Australia and New Zealand had
experienced warming of 0.4°C to 0.7°C since 1950 with changed
rainfall patterns and sea level rise of about 70 mm across the region;
there had also been a greater frequency and intensity of droughts
and heat waves, reduced seasonal snow cover, and glacial retreat.
Impacts from recent climate changes were evident in increasing
stresses on water supply and agriculture, and changed natural
ecosystems; some adaptation had occurred in these sectors but
vulnerability to extreme events such as fire, tropical cyclones,
droughts, hail, and floods remained high.
The climate of the 21st century would be warmer (virtually certain),
with changes in extreme events including more intense and frequent
heat waves, fire, floods, storm surges, and droughts but less frequent
frost and snow (high confidence), reduced soil moisture in large
parts of the Australian mainland and eastern New Zealand but more
rain in western New Zealand (medium confidence).
Significant advances had occurred in understanding future impacts
on water, ecosystems, indigenous people and health, together with
an increased focus on adaptation; potential impacts would be
substantial without further adaptation, particularly for water security,
coastal development, biodiversity, and major infrastructure, but
impacts on agriculture and forestry would be variable across the
region, including potential benefits in some areas.
Vulnerability would increase mainly due to an increase in extreme
events; human systems were considered to have a higher adaptive
capacity than natural systems.
Hotspots of high vulnerability by 2050 under a medium emissions
scenario included:
Significant loss of biodiversity in areas such as alpine regions,
the Wet Tropics, the Australian southwest, Kakadu wetlands,
coral reefs, and sub-Antarctic islands
Water security problems in the Murray-Darling basin,
southwestern Australia, and eastern New Zealand
Potentially large risks to coastal development in southeastern
Queensland and in New Zealand from Northland to the Bay of
Plenty.
25.2. Observed and Projected Climate Change
Australasia exhibits a wide diversity of climates, such as moist tropical
monsoonal, arid, and moist temperate, including alpine conditions. Key
climatic processes are the Asian-Australian monsoon and the southeast
trade winds over northern Australia, and the subtropical high pressure
b
elt and the mid-latitude storm tracks over southern Australia and New
Zealand. Tropical cyclones also affect northern Australia, and, more
rarely, ex-tropical cyclones affect some parts of New Zealand. Natural
climatic variability is very high in the region, especially for rainfall and
over Australia, with the El Niño-Southern Oscillation (ENSO) being the
most important driver (McBride and Nicholls, 1983; Power et al., 1998;
Risbey et al., 2009). The southern annular mode, Indian Ocean Dipole,
and the Inter-decadal Pacific Oscillation are also important regional
drivers (Thompson and Wallace, 2000; Salinger et al., 2001; Cai et al.,
2009b). This variability poses particular challenges for detecting and
projecting anthropogenic climate change and its impacts in the region.
For example, changes in ENSO in response to anthropogenic climate
change are uncertain (WGI AR5 Chapter 14) but, given current ENSO
impacts, any changes would have the potential to significantly influence
rainfall and temperature extremes, droughts, tropical cyclones, marine
conditions, and glacial mass balance (Mullan, 1995; Chinn et al., 2005;
Holbrook et al., 2009; Diamond et al., 2012; Min et al., 2013).
Understanding of observed and projected climate change has received
much attention since AR4, particularly in Australia, with a focus on the
causes of observed rainfall changes and more systematic analysis of
projected changes from different models and approaches. Climatic
extremes have also been a research focus. Table 25-1 presents an
assessment of this body of research for observed trends and projected
changes for a range of climatic variables (including extremes) relevant
for regional impacts and adaptation, including examples of the magnitude
of projected change, and attribution, where possible. Most studies are
based on Coupled Model Intercomparison Project Phase 3 (CMIP3)
models and Special Report on Emission Scenarios (SRES) scenarios, but
CMIP5 model results are considered where available (see also WGI AR5
Chapter 14 and Atlas; Chapter 21).
The region has exhibited warming to the present (very high confidence)
and is virtually certain to continue to do so (Table 25-1). Observed and
CMIP5-modeled past and projected future annual average surface
temperatures are shown in Figures 25-1 and 25-2. For further details
see WGI AR5 Atlas, AI.68–69. Changes in precipitation have been
observed with very high confidence in some areas over a range of time
scales, such as increases in northwestern Australia since the 1950s, the
autumn/winter decline since 1970 in southwestern Australia, and, since
the 1990s, in southeastern Australia, and over 1950–2004 increases in
annual rainfall in the south and west of the South Island and west of
the North Island of New Zealand, and decreases in the northeast of the
South Island and east and north of the North Island. Based on multiple
lines of evidence, annual average rainfall is projected to decrease with
high confidence in southwestern Australia. For New Zealand, annual
average rainfall is projected to decrease in the northeastern South Island
and eastern and northern North Island, and increase in other parts of
the country (medium confidence). The direction and magnitude of rainfall
change in eastern and northern Australia remains a key uncertainty
(Table 25-1).
This pattern of projected rainfall change is reflected in annual average
CMIP5 model results (Figure 25-1), but with important additional
dimensions relating to seasonal changes and spread across models (see
also WGI AR5 Atlas, AI.70–71). Examples of the magnitude of projected
annual change from 1990 to 2090 (percent model mean change +/–
1378
Chapter 25 Australasia
25
Annual Precipitation
Change
Diagonal Lines
Trend not
statistically
significant
White
Insufficient
data
Solid Color
Strong
agreement
Very strong
agreement
Little or
no change
Gray
Divergent
changes
Solid Color
Significant
trend
Diagonal Lines
White Dots
Annual Temperature Change
late-21st century
mid-21st century
Difference from 1986–2005 mean (%)
Difference from 1986–2005 mean
(˚C)
Trend over 19012012
(˚C over period)
0
246
(mm/year per decade)
Trend in annual precipitation over 1951–2010
5 0525102.52.5 501050 25100
Figure 25-1 | Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Map of observed annual average temperature change from
1901–2012, derived from a linear trend. [WGI AR5 Figures SPM.1 and 2.21] (Bottom panel, left) Map of observed annual precipitation change from 1951–2010, derived from a
linear trend. [WGI AR5 Figures SPM.2 and 2.29] For observed temperature and precipitation, trends have been calculated where sufficient data permit a robust estimate (i.e., only
for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors
indicate areas where trends are significant at the 10% level. Diagonal lines indicate areas where trends are not significant. (Top and bottom panel, right) CMIP5 multi-model
mean projections of annual average temperature changes and average percent changes in annual mean precipitation for 2046–2065 and 2081–2100 under RCP2.6 and 8.5,
relative to 1986–2005. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability (natural internal
variability in 20-yr means) and ≥90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where ≥66% of models show change
greater than the baseline variability and ≥66% of models agree on sign of change. Gray indicates areas with divergent changes, where ≥66% of models show change greater
than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, where <66% of models show change greater
than the baseline variability, although there may be significant change at shorter timescales such as seasons, months, or days. Analysis uses model data and methods building
from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5. [Boxes 21-2 and CC-RC]
–20 0 20 40
RCP8.5RCP2.6
late 21st century
mid 21st century
RCP8.5RCP2.6
1379
Australasia Chapter 25
25
i
ntermodel standard deviation) under Representative Concentration
Pathway (RCP)8.5 from CMIP5 are –20 ± 13% in southwestern
Australia, –2 ± 21% in the Murray-Darling Basin, and –5 ± 22% in
southeast Queensland (Irving et al., 2012). Projected changes during
winter and spring are more pronounced and/or consistent across models
than the annual changes, for example, drying in southwestern Australia
(–32 ± 11%, June to August), the Murray-Darling Basin (–16 ± 22%,
June to August), and southeast Queensland (–15 ± 26%, September to
November), whereas there are increases of 15% or more in the west
and south of the South Island of New Zealand (Irving et al., 2012).
Downscaled CMIP3 model projections for New Zealand indicate a
stronger drying pattern in the southeast of the South Island and eastern
and northern regions of the North Island in winter and spring (Reisinger
et al., 2010) than seen in the raw CMIP5 data; based on similar broader
scale changes this pattern is expected to hold once CMIP5 data are also
downscaled (Irving et al., 2012).
O
ther projected changes of at least high confidence include regional
increases in sea surface temperature, the occurrence of hot days, fire
weather in southern Australia, mean and extreme sea level, and ocean
acidity (see WGI AR5 Section 6.4.4 for projections); and decreases in
cold days and snow extent and depth. Although changes to tropical
cyclone occurrence and that of other severe storms are potentially
important for future vulnerability, regional changes to these phenomena
cannot be projected with at least medium confidence as yet (Table 25-1).
25.3. Socioeconomic Trends Influencing
Vulnerability and Adaptive Capacity
25.3.1. Economic, Demographic, and Social Trends
The economies of Australia and New Zealand rely on natural resources,
agriculture, minerals, manufacturing and tourism, but the relative
importance of these sectors differs between the two countries.
Agriculture and mineral/energy resources accounted, respectively, for 11%
and 55% (Australia) and 56% and 5% (New Zealand) of the value of total
exports in 2010–2011 (ABS, 2012c; SNZ, 2012b). Water abstraction per
capita in both countries is in the top half of the Organisation for
Economic Co-operation and Development (OECD), decreasing since
1990 in Australia but increasing in New Zealand; more than half is used
for irrigation (OECD, 2010, 2013a). Between 1970 and 2011, gross
domestic product (GDP) grew by an average of 3.2% per annum in
Australia and 2.4% per annum in New Zealand, with annual GDP per
capita growth of 1.8% and 1.2%, respectively (SNZ, 2011; ABS, 2012d).
GDP is projected to grow on average by 2.5 to 3.5% per annum in
Australia and about 1.9% per annum in New Zealand to 2050 (Australian
Treasury, 2010; Bell et al., 2010) but subject to significant shorter term
fluctuations.
The populations of Australia and New Zealand are projected to grow
significantly over at least the next several decades (very high confidence;
ABS, 2008; SNZ, 2012a): Australia’s population from 22.3 million in 2011
to 31 to 43 million by 2056 and 34 to 62 million by 2101 (ABS, 2008,
2013); New Zealand’s population from 4.4 million in 2011 to 5.1 to 7.1
million by 2061 (SNZ, 2012a). The number of people aged 65 and over
is projected to almost double in the next 2 decades (ABS, 2008; SNZ,
2012a). More than 85% of the Australasian population lives in urban
areas and their satellite communities, mostly in coastal areas (DCC,
2009; SNZ, 2010b; UN DESA Population Division, 2012; see Box 25-9).
Urban concentration and depletion of remote rural areas is expected to
continue (Mendham and Curtis, 2010; SNZ, 2010c; Box 25-5), but some
coastal non-urban spaces also face increasing development pressure
(Freeman and Cheyne, 2008; Gurran, 2008; Box 25-1). More than 20%
of Australasian residents were born overseas (OECD, 2013a).
Poverty rates and income inequality in Australia and New Zealand are
in the upper half of OECD countries, and both measures increased
significantly in both countries between the mid-1980s and the late
2000s (OECD, 2013a). Measurement of poverty and inequality, however,
is highly contested, and it remains difficult to anticipate future changes
and their effects on adaptive capacity (Peace, 2001; Scutella et al., 2009;
Section 25.3.2). Indigenous peoples constitute about 2.5% and 15% of
the Australian and New Zealand populations, respectively, but in
Historical
Natural
Overlap
Observed
1900 1950 2000 2050 2100
–2
0
2
4
6
8
°C relative to 19862005
New Zealand
1900 1950 2000 2050 2100
–2
0
2
4
6
8
°C relative to 19862005
Australia
Figure 25-2 | Observed and simulated variations in past and projected future annual
average near-surface air temperature over land areas of Australia (top) and New
Zealand (bottom). Black lines show various estimates from observational measure-
ments. Shading denotes the 5th to 95th percentile range of climate model simulations
driven with “historical” changes in anthropogenic and natural drivers (63 simulations),
historical changes in “natural drivers only (34), the Representative Concentration
Pathway (RCP)2.6 emissions scenario (63), and the RCP8.5 (63). Data are anomalies
from the 1986–2005 average of the individual observational data (for the observa-
tional time series) or of the corresponding historical all-forcing simulations. Further
details are given in Box 21-3 and Box CC-RC.
Near-surface air temperature
Overlap
RCP2.6
RCP8.5
1380
Chapter 25 Australasia
25
Climate
variable
Observed change Direction of projected change
Examples of projected magnitude of change
(relative to ~1990, unless otherwise stated)
Additional comments
Mean air
temperature
Australia: Increased by 0.09 ± 0.03°C per decade since 1911
1
(***)
New Zealand: Increased by 0.09 ± 0.03°C per decade since
1909
2
(***)
Australia and New Zealand: Increase
3–8
(****);
greatest over inland Australia and least in coastal
areas and New Zealand
5–8
(***)
Australia: 0.6 1.5°C (2030 A1B), 1.0 2.5°C (2070 B1),
2.2 5.0°C (2070 A1FI)
3
New Zealand: 0.3 1.4°C (2040 A1B), 0.7 2.3°C (2090 B1),
1.6 5.1°C (2090 A1FI)
5
Coupled Model Intercomparison Project Phase 5 (CMIP5)
Representative Concentration Pathway 4.5 (RCP4.5), relative
to ~1995
9
:
North Australia: 0.3 1.6°C (2016 2035), 0.7 2.6°C
(2046 2065)
Southern Australia and New Zealand: 0.1 1.0°C
(2016 2035), 0.6 1.7°C (2046 2065)
Australia: A signifi cant contribution to observed change
attributed to anthropogenic climate change
10
(**) with
some regional variations attributed to atmospheric
circulation variations
11,12
New Zealand: Observed change partially attributed to
anthropogenic climate change
13
(*)
Sea surface
temperature
Australia: Increased by about 0.12°C per decade for
northwestern and northeastern Australia and by about 0.2°C
per decade for southeastern Australia since 1950
14,15
(***)
New Zealand: Increased by about 0.07°C per decade over
1909 2009
2
(***)
Australia and New Zealand: Increase
3,7,8
(***), with
greater increase in the Tasman sea region
3,7
(*)
Australia: 0.6 1.0°C (2070 B1) and 1.6 2.0°C (2070
A1FI) for southern coastal and 1.2 1.5°C (2070 B1) and
2.2 2.5°C (2070 A1FI) elsewhere
3
New Zealand: Similar to projected changes in mean air
temperature for coastal waters
5
Air
temperature
extremes
Australia and New Zealand: Signifi cant trend since 1950: Cool
extremes have become rarer and hot extremes more frequent
and intense
16–19
(**). The Australian heat wave of 2012 / 13 was
exceptional in heat, duration, and spatial extent.
20
Australia and New Zealand: Hot days and nights
more frequent and cold days and cold nights less
frequent during the 21st century
3,5,21–24
(**)
Australia: Hot days in Melbourne (>35°C max.) increase by
20 40% (2030 A1B), 30 90% (2070 B1), and 70 190%
(2070 A1FI)
3
New Zealand: Spring and autumn frost–free land to at least
triple by 2080s
24
; up to 60 more hot days (>25°C max.) for
northern areas by 2090
5
Australia: Observed trends partly attributable to
anthropogenic climate change (**) as they are consistent
with mean warming and historical simulations,
18,19,21,25
although other factors may have contributed to high
extremes during droughts
26–28
Precipitation
Australia: Late autumn / winter decreases in southwestern
Australia since the 1970s and in southeastern Australia since
the mid–1990s, and annual increases in northwestern Australia
since the 1950s
29–31
(***)
New Zealand: Mean annual rainfall increased over 1950 2004
in the south and west of the South Island and west of the
North Island, and decreased in the northeast of the South
Island and east and north of the North Island
32
(***).
Australia: Annual decline in southwestern
Australia (**), elsewhere on most of the southern
(*) and northeastern ( low confi dence) continental
edges, with reductions strongest in the winter
half year
3,4,9,33–35
(**). Direction of annual change
elsewhere is uncertain
3,35,36
(Figure 25-1) (**).
New Zealand: In the South Island, annual
increase in the west and south and decrease
in northeast. In the North Island, increase in
the west and decrease in eastern and northern
regions
5,34,37
(Figure 25-1) (*)
Australia: For 2030 A1B, annual changes of –10% to +5%
(northern Australia) and –10% to 0% (southern Australia);
for 2070 B1, –15% to +7.5% (northern and eastern
Australia) and –15% to 0% (southern Australia); and for
2070 A1FI, –30% to +20% (northern and eastern Australia)
and –30% to +5% (southern Australia), with larger changes
seasonally
3
New Zealand: For 2040 A1B, annual changes of –5%
to +15% (southern and western) and –15% to +10%
(northern and eastern) and for 2090 A1B, –10% to +25%
(southern and western) and –20% to +15% (northern
and eastern) based on downscaled projections with larger
changes seasonally
5,37
Australia: Observed decline in southwest is related to
atmospheric circulation changes
38–40
(***) and other
factors,
41
and partly attributable to anthropogenic climate
change
40–43
(**). The recent southeast rainfall decline is
also related to circulation changes
31,44–46
(**), with some
evidence of an anthropogenic component.
47
New Zealand: Observed trends related to increased
westerly winds.
32
Projected annual trends dominated by
winter and spring trends related to increased westerlies
5
Precipitation
extremes
Australia: Indices of annual daily extremes (e.g., 95th and 99th
percentile rainfalls) show mixed or insignifi cant trends,
21,48
but
signifi cant increase is evident in recent decades for shorter
duration (sub-daily) events
49,50
(**).
New Zealand: Extreme annual 1-day rainfall decrease in north
and east and increase in west since 1930
32
(*)
Australia and New Zealand: Increase in most
regions in the intensity of rare daily rainfall
extremes (i.e., current 20-year return period
events) and in short duration (sub-daily)
extremes (*) and an increase in the intensity
of 99th percentile daily extremes ( low
confi dence)
5,8,21,51–56
Australia: For 2090 A2, CMIP3 gives increases in the
intensity of the 20-year daily extreme of around +200% to
–25% depending on region and model.
52
New Zealand: Increases of daily extreme rainfalls of around
8% per degree Celsius are projected but with signifi cant
regional variations.
5,56
Australia and New Zealand: The sign of observed trends
mostly refl ects trends in mean rainfall (e.g., there is a
decrease in mean and daily extremes in southwestern
Australia).
21,32,49
Similarly, future increases in intensity of
extreme daily rainfall are more likely where mean rainfall
is projected to increase.
3,5
Drought
Australia: Defi ned using rainfall only, drought occurrence over
the period 1900 2007 has not changed signifi cantly
57
(**).
New Zealand: Defi ned using a soil water balance model; there
has been no trend in drought occurrence since 1972
58
(*).
Australia and New Zealand: Drought
frequency is projected to increase in southern
Australia
8,54,57,59,60
(*) and in many regions of New
Zealand
58,61
(*).
Australia: Occurrence under 2070 A1B and A2 ranges from a
halving to 3 times more frequent in northern Australia and
0 5 times more frequent in southern Australia.
60
New Zealand: Time spent in drought in eastern and northern
New Zealand is projected to double or triple by 2040.
61
Australia: Regional warming may have led to an increase
in hydrological drought ( low confi dence).
62,63
Table 25-1 | Observed and projected changes in key climate variables, and (where assessed) the contribution of human activities to observed changes. For further relevant information see WGI AR5 Chapters 3, 6 (ocean changes, including
acidifi cation), 11, 12 (projections), 13 (sea level), and 14 (regional climate phenomena). (*) medium confi dence, (**) high confi dence, (***) very high confi dence, (****) virtually certain
1381
Australasia Chapter 25
25
1382
Chapter 25 Australasia
25
A
ustralia, their national share is growing and they constitute a much
higher percentage of the population in remote and very remote regions
(ABS, 2009, 2010b; SNZ, 2010a). Indigenous peoples in both countries
have lower than average life expectancy, income, and education, implying
that changes in socioeconomic status and social inclusion could strongly
influence their future adaptive capacity (see Section 25.8.2).
25.3.2. Use and Relevance of Socioeconomic Scenarios in
Adaptive Capacity/Vulnerability Assessments
Demographic, economic, and sociocultural trends influence the
vulnerability and adaptive capacity of individuals and communities
(see Chapters 2, 11-13, 16, 20). A limited but growing number of studies
in Australasia have attempted to incorporate such information, for
example, changes in the number of people and percentage of elderly
people at risk (Preston et al., 2008; Baum et al., 2009; Preston and
Stafford-Smith, 2009; Roiko et al., 2012), the density of urban settlements
and exposed infrastructure (Preston and Jones, 2008; Preston et al.,
2008; Baynes et al., 2012), population-driven pressures on water demand
(Jollands et al., 2007; CSIRO, 2009), and economic and social factors
affecting individual coping, planning, and recovery capacity (Dwyer et
al., 2004; Khan, 2012; Roiko et al., 2012).
Socioeconomic considerations are used increasingly to understand
adaptive capacity of communities (Preston et al., 2008; Smith et al.,
2008; Fitzsimons et al., 2010; Soste, 2010; Brunckhorst et al., 2011) and
to construct scenarios to help build regional planning capacity (Energy
Futures Forum, 2006; Frame et al., 2007; Pride et al., 2010; Pettit et al.,
2011; Taylor et al., 2011). Such scenarios, however, are only beginning
to be used to quantify vulnerability to climate change (except, e.g.,
Bohensky et al., 2011; Baynes et al., 2012; Low Choy et al., 2012).
Apart from these emerging efforts, most vulnerability studies from
Australasia make no or very limited use of socioeconomic factors,
consider only current conditions, and/or rely on postulated correlations
between generic socioeconomic indicators and climate change
vulnerability. In many cases this limits confidence in conclusions
regarding future vulnerability to climate change and adaptive capacity
of human and mixed natural-human systems.
25.4. Cross-Sectoral Adaptation:
Approaches, Effectiveness, and Constraints
25.4.1. Frameworks, Governance,
and Institutional Arrangements
Adaptation responses depend heavily on institutional and governance
arrangements (see Chapters 2, 14-16, 20). Responsibility for development
and implementation of adaptation policy in Australasia is largely
devolved to local governments and, in Australia, to State governments
and Natural Resource Management bodies. Federal/central government
supports adaptation mostly via provision of information, tools, legislation,
policy guidance, and (in Australia) support for pilot projects. A standard
risk management paradigm has been promoted to embed adaptation
into decision-making practices (AGO, 2006; MfE, 2008b; Standards
A
ustralia, 2013), but broader systems and resilience approaches are used
increasingly for natural resource management (Clayton et al., 2011;
NRC, 2012). The Council of Australian Governments agreed a national
adaptation policy framework in 2007 (COAG, 2007). This included
establishing the collaborative National Climate Change Adaptation
Research Facility (NCCARF) in 2008, which complemented Commonwealth
Scientific and Industrial Research Organisation (CSIRO)’s Climate
Adaptation Flagship. The federal government supported a first-pass
national coastal risk assessment (DCC, 2009; DCCEE, 2011), is developing
indicators and criteria for assessing adaptation progress and outcomes
(DIICCSRTE, 2013), and commissioned targeted reports addressing
impacts and management options for natural and managed landscapes
(Campbell, 2008; Steffen et al., 2009; Dunlop et al., 2012), National and
World Heritage areas (ANU, 2009; BMT WBM, 2011), and indigenous
and urban communities (Green et al., 2009; Norman, 2010). Most State
and Territory governments have also developed adaptation plans (e.g.,
DSE, 2013).
In New Zealand, the central government updated and expanded tools
to support impact assessments and adaptation responses consistent
with regulatory requirements (MfE, 2008b,c,d, 2010b), and revised key
directions for coastal management (Minister of Conservation, 2010).
No cross-sectoral adaptation policy framework or national-level risk
assessments exist, but some departments commissioned high-level
impacts and adaptation assessments after the AR4 (e.g., on agriculture
and on biodiversity; Wratt et al., 2008; McGlone and Walker, 2011; Clark
et al., 2012).
Public and private sector organizations are potentially important
adaptation actors but exhibit large differences in preparedness, linked
to knowledge about climate change, economic opportunities, external
connections, size, and scope for strategic planning (Gardner et al., 2010;
Taylor, B.M. et al., 2012; Johnston et al., 2013; Kuruppu et al., 2013; see
also Chapters 10, 16). This creates challenges for achieving holistic
societal outcomes (see also Sections 25.7-9).
Several recent policy initiatives in Australia, while responding to broader
socioeconomic and environmental pressures, include goals to reduce
vulnerability to climate variability and change. These include establishing
the Murray-Darling Basin Authority to address over-allocation of water
resources (Connell and Grafton, 2011; MDBA, 2011), removal of the
interest rate subsidy during exceptional droughts (Productivity
Commission, 2009), and management of bush fire and flood risk (VBRC,
2010; QFCI, 2012). These may be seen as examples of mainstreaming
adaptation (Dovers, 2009), but they also demonstrate lag times in policy
design and implementation, windows of opportunity presented by crises
(e.g., the Millennium Drought of 1997–2009, the Victorian bushfires of
2009, and Queensland floods of 2011), and the challenges arising from
competing interests in managing finite and changing water resources
(Botterill and Dovers, 2013; Pittock, 2013; Box 25-2).
25.4.2. Constraints on Adaptation and
Emerging Leading Practice Models
A rapidly growing literature since the AR4 confirms, with high confidence,
that while the adaptive capacity of society in Australasia is generally high,
1383
Australasia Chapter 25
25
there are formidable environmental, economic, informational, social,
attitudinal, and political constraints, especially for local governments
and small or highly fragmented industries. Reviews of public- and private-
sector adaptation plans and strategies in Australia demonstrate strong
efforts in institutional capacity building, but differences in assessment
methods and weaknesses in translating goals into specific policies
(White, 2009; Gardner et al., 2010; Measham et al., 2011; Preston et al.,
2011; Kay et al., 2013). Similarly, local governments in New Zealand to
date have focused mostly on impacts and climate-related hazards; some
have developed adaptation plans, but few have committed to specific
policies and steps to implementation (e.g., O’Donnell, 2007; Britton,
2010; Fitzharris, 2010; HRC, 2010; KCDC, 2012; Lawrence et al., 2013b).
Table 25-2 summarizes key constraints and corresponding enabling factors
for effective institutional adaptation processes identified in Australia
and New Zealand. Scientific uncertainty and resource limitations are
reported consistently as important constraints, particularly for smaller
councils. Ultimately more powerful constraints arise, however, from
current governance and legislative arrangements and the lack of
consistent tools to deal with dynamic risks and uncertainty or to evaluate
the success of adaptation responses (robust evidence, high agreement;
Britton, 2010; Barnett et al., 2013; Lawrence et al., 2013b; Mukheibir
et al., 2013; Webb et al., 2013; see also Chapter 16).
Some constraints exacerbate others. There is high confidence that the
absence of a consistent information base and binding guidelines that
clarify governing principles and liabilities is a challenge particularly for
small and resource-limited local authorities, which need to balance
special interest advocacy with longer term community resilience. This
heightens reliance on individual leadership subject to short-term
political change and can result in piecemeal and inconsistent risk
assessments and responses between levels of government and locations,
and over time (Smith et al., 2008; Brown et al., 2009; Norman, 2009;
Britton, 2010; Rouse and Norton, 2010; Abel et al., 2011; McDonald,
2011; Rive and Weeks, 2011; Corkhill, 2013; Macintosh et al., 2013). In
these situations, planners tend to rely more on single numbers for
climate projections that can be argued in court (Reisinger et al., 2011;
Lawrence et al., 2013b), which increases the risk of maladaptation given
Constraint Enabling factors
U
ncertainty of projections
I
mproved guidance and tools to manage uncertainty and support adaptive management
1 8
Increased focus on lead and consequence time of decisions and link with current climate variability and related risks
9
13
Increased communication between practitioners and scientists to identify and provide decision-relevant data and context
2,3,11,13 17
A
vailability and cost of data and models
C
entral provision of relevant core climate and non-climate data, including regional scenarios of projected changes
4
,5,7,9,18,19
National fi rst-pass risk assessments
4,5,7,8,18,20 24
L
imited fi nancial and human capability and
capacity; time lag in developing expertise
S
upport for pilot projects
4,8,15,18,24,25
B
uilding capacity through institutional commitment and learning
3,5,11,17,23,26 28
Central databases on guidance, tools, methodologies, case studies
4
,5,7,18,24
Regional partnerships and collaborations, knowledge networks
3,4,8,13,15,17,26,28 30
Unclear problem defi nition and goals;
u
nclear standards for risk assessment
methodologies and decision support tools;
l
imited monitoring and evaluation
Explicit but iterative framing and scoping of adaptation challenge, to refl ect alternative entry points for stakeholders while meeting
e
xpectations of project sponsors to ensure long-term support
3
,11,17,31– 34
T
ailoring decision-making frameworks to specifi c problems
1,2,6,17,35,36
Criteria and tools to monitor and evaluate adaptation success
7,18,37– 39
U
nclear or contradictory legislative
frameworks and responsibilities, unclear
l
iabilities
C
lear and coordinated legislative frameworks
5,8,9,15,24,40 45
Defi ned responsibilities for public and private actors, including liabilities from acting and failure to act
8,9,11,24,41,44,46
Legally binding guidance on the incorporation of climate change in planning mechanisms
5
,7,8,15,38,40
Static planning mechanisms and practice;
competing mandates and fragmentation
o
f policies; disciplinary voids or single
approaches
Whole-of-council approach to climate adaptation to break up institutional and professional silos
1
5,33,47
L
ong-term policy commitments and implementation support
5,18,26,33,48
Increased policy coherence across sectors, regulations, and levels of government
9,26,28,40,42,43,47
Enabling risk-based fl exible land use decisions
4
,5,9,49
S
trengthening multi-disciplinarity across professional fi elds
14,29,48
Lack of political leadership; short election
cycles; limited community support,
participation, and awareness for adaptation
Legally binding guidance and clarifi cation of liabilities and duty of care to reduce dependence on individual leadership
5 , 7 9 , 1 5 , 2 4 , 3 8 , 4 0 , 4 6 , 4 9
Consistent but audience-speci c communication of current and potential future vulnerability and implications for community values
4
,5,7,26,42,43,50
Comprehensible communication of and access to response options, and their consistency with wider development plans
7,26,28,33,39,42,43
Clearly identifi ed entry points for public participation
17,34,38,39,42,48,51– 53
Table 25-2 | Constraints and enabling factors for institutional adaptation processes in Australasia.
Note: The relevance of each constraint varies among organizations, sectors, and locations. Some enabling factors are only beginning to be implemented or have only been
suggested in the literature; hence their effectiveness cannot yet be evaluated. Entries for enabling factors exclude generic mechanisms, such as insurance (see Box 25-7);
emergency management and early warning systems; and funding for pilot studies, capital infrastructure upgrades, or retreat schemes.
References:
1
Randall et al. (2012);
2
Verdon-Kidd et al. (2012);
3
Webb et al. (2013);
4
Mukheibir et al. (2013);
5
Lawrence et al. (2013b);
6
Nelson et al. (2008);
7
Britton (2010);
8
Gurran et al. (2008);
9
Productivity Commission (2012);
10
Stafford-Smith et al. (2011);
11
Johnston et al. (2013);
12
Park et al. (2012);
13
Power et al. (2005);
14
Reisinger et al. (2011);
15
Smith et al. (2008);
16
Stafford-Smith (2013);
17
Yuen et al. (2012);
18
Webb and Beh (2013);
19
Roiko et al. (2012);
20
DCCEE (2011);
21
DCC (2009);
22
Baynes et al. (2012);
23
Smith
et al. (2010);
24
SCCCWEA (2009);
25
DSEWPC (2011);
26
Low Choy et al. (2012);
27
Gardner et al. (2010);
28
Fidelman et al. (2013);
29
Mustelin et al. (2013);
30
Serrao-Neumann et
al. (2013);
31
Fünfgeld et al. (2012);
32
Kuruppu et al. (2013);
33
Britton et al. (2011);
34
Alexander et al. (2012);
35
Maru et al. (2011);
36
Preston et al. (2008);
37
Norman et al. (2013);
38
Rouse and Norton (2010);
39
Preston et al. (2011);
40
Rive and Weeks (2011);
41
Abel et al. (2011);
42
Norman (2009);
43
Gurran et al. (2006);
44
McDonald (2013);
45
Minister of
Conservation (2010);
46
McDonald (2010);
47
Measham et al. (2011);
48
Rouse and Blackett (2011);
49
McDonald (2011);
50
Hine et al. (2013);
51
Burton and Mustelin (2013);
52
Hobson
and Niemeyer (2011);
53
Gardner et al. (2009a).
1384
Chapter 25 Australasia
25
Box 25-1 | Coastal Adaptation—Planning and Legal Dimensions
Sea level rise is a significant risk for Australia and New Zealand (very high confidence) due to intensifying coastal development and
the location of population centers and infrastructure (see Section 25.3). Under a high emissions scenario (Representative Concentration
Pathway (RCP)8.5), global mean sea level would likely rise by 0.53 to 0.97 m by 2100, relative to 1986–2005, whereas with stringent
mitigation (RCP2.6), the likely rise by 2100 would be 0.28 to 0.6 m (medium confidence). Based on current understanding, only
instability of the Antarctic ice sheet, if initiated, could lead to a rise substantially above the likely range; evidence remains insufficient
to evaluate its probability, but there is medium confidence that this additional contribution would not exceed several tenths of a
meter during the 21st century (WGI AR5 Section 13.5). Local case studies in New Zealand (Fitzharris, 2010; Reisinger et al., 2013) and
national reviews in Australia (DCC, 2009; DCCEE, 2011) demonstrate risks to large numbers of residential and commercial assets as
well as key services, with widespread damages at the upper end of projected ranges (high confidence). In Australia, sea level rise of
1.1 m would affect more than AU$226 billion of assets, including up to 274,000 residential and 8600 commercial buildings (DCCEE,
2011), with additional intangible costs related to stress, health effects, and service disruption (HCCREMS, 2010) and ecosystems
(DCC, 2009; BMT WBM, 2011). Under expected future settlement patterns, exposure of the Australian road and rail network will
increase significantly once sea level rises above about 0.5 m (Baynes et al., 2012). Even if temperatures peak and decline, sea level is
projected to continue to rise beyond 2100 for many centuries, at a rate dependent on future emissions (WGI AR5 Section 13.5).
Responsibility for adapting to sea level rise in Australasia rests principally with local governments through spatial planning instruments.
Western Australia, South Australia, and Victoria have mandatory State planning benchmarks for 2100, with local governments
determining how they should be implemented. Long-term benchmarks in New South Wales and Queensland have either been
suspended or revoked, so local authorities now have broad discretion to develop their own adaptation plans. The New Zealand
Coastal Policy Statement (Minister of Conservation, 2010) mandates a minimum 100-year planning horizon for assessing hazard
risks, discourages hard protection of existing development, and recommends avoidance of new development in vulnerable areas.
Non-binding government guidance recommends a risk-based approach, using a base value of 0.5 m sea level rise by the 2090s and
considering the implications of at least 0.8 m and, for longer term planning, an additional 0.1 m per decade (MfE, 2008d).
The incorporation of climate change impacts into local planning has evolved considerably over the past 20 years, but remains piecemeal
and shows a diversity of approaches (Gibbs and Hill, 2012; Kay et al., 2013). Governments have invested in high-resolution digital
elevation models of coastal and flood prone areas in some regions, but many local governments still lack the resources for hazard
mapping and policy design. Political commitment is variable, and legitimacy of approaches and institutions is often strongly
contested (Gorddard et al., 2012), including pressure on State governments to modify adaptation policies and on local authorities to
compensate developers for restrictions on current or future land uses (LGNZ, 2008; Berry and Vella, 2010; McDonald, 2010; Reisinger
et al., 2011). Incremental adaptation responses can entrench existing rights and expectations about ongoing protection and
development, which limit options for more transformational responses such as accommodation and retreat (medium evidence, high
agreement; Gorddard et al., 2012; Barnett et al., 2013; Fletcher et al., 2013; McDonald, 2013). Strategic regional-scale planning
initiatives in rapidly growing regions, like southeast Queensland, allow climate change adaptation to be addressed in ways not
typically achieved by locality- or sector-specific plans, but require effective coordination across different scales of governance (Serrao-
Neumann et al., 2013; Smith et al., 2013).
Courts in both countries have played an important role in evaluating planning measures. Results of litigation have varied and, in the
absence of clearer legislative guidance, more litigation is expected as rising sea levels affect existing properties and adaptation
responses constrain development on coastal land (MfE, 2008d; Kenderdine, 2010; Rive and Weeks, 2011; Verschuuren and McDonald,
2012; Corkhill, 2013; Macintosh, 2013).
Continued next page
1385
Australasia Chapter 25
25
the uncertain and dynamic nature of climate risk (McDonald, 2010;
Stafford-Smith et al., 2011; Gorddard et al., 2012; McDonald, 2013;
Reisinger et al., 2013).
Vulnerability assessments that take mid- to late-century impacts as their
starting point can inhibit actors from implementing adaptation actions,
as distant impacts are easily discounted and difficult to prioritize in
competition with near-term non-climate change pressures (Productivity
Commission, 2012). Emerging leading practice models in Australia
(Balston, 2012; HCCREMS, 2012; SGS, 2012) and New Zealand (MfE,
2008a; Britton et al., 2011) recommend a high-level scan of sectors and
locations at risk and emphasize a focus on near-term decisions that
influence current and future vulnerability (which could range from early
warning systems to strategic and planning responses). More detailed
assessment can then focus on this more tractable subset of issues, based
on explicit and iterative framing of the adaptation issue (Webb et al.,
2013) and taking into account the full lifetime (lead- and consequence
time) of the decision/asset in question (Stafford-Smith et al., 2011).
Participatory processes help balance societal preferences with robust
scientific information and ensure ownership by affected communities
but rely on human capital and political commitment (high confidence;
Hobson and Niemeyer, 2011; Rouse and Blackett, 2011; Weber et al.,
2011; Leitch and Robinson, 2012). Realizing widespread and equitable
participation is challenging where policies are complex, debates polarized,
legitimacy of institutions contested, and potential transformational
changes threaten deeply held values (Gardner et al., 2009a; Gorddard
et al., 2012; Burton and Mustelin, 2013; see also Section 25.4.3).
Regional approaches that engage diverse stakeholders, government, and
science providers, and support the co-production of knowledge can help
overcome some of these problems but require long-term institutional
and financial commitments (e.g., Britton et al., 2011; DSEWPC, 2011;
CSIRO, 2012; IOCI, 2012; Low Choy et al., 2012; Webb and Beh, 2013).
There is active debate about the extent to which incremental adjustments
of existing planning instruments, institutions, and decision-making
processes can deal adequately with the dynamic and uncertain nature
of climate change and support transformational responses (Kennedy et
al., 2010; Preston et al., 2011; Park et al., 2012; Dovers, 2013; Lawrence
et al., 2013b; McDonald, 2013; Stafford-Smith, 2013). Recent studies
suggest a greater focus on flexibility and matching decision-making
frameworks to specific problems (Hertzler, 2007; Nelson et al., 2008;
Dobes, 2010; Howden and Stokes, 2010; Randall et al., 2012). Limitations
of mainstreamed and autonomous adaptation and the case for more
proactive government intervention are being explored in Australia
(Productivity Commission, 2012; Johnston et al., 2013), but have not
yet resulted in new policy frameworks.
25.4.3. Psychological and Sociocultural Factors Influencing
Impacts of and Adaptation to Climate Change
Adapting to climate change relies on individuals accepting and
understanding changing risks and opportunities, and responding to
these changes both psychologically and behaviorally (see Chapters 2,
16). The majority of Australasians accept the reality of climate change
and less than 10% fundamentally deny its existence (high confidence;
ShapeNZ, 2009; Leviston et al., 2011; Lewandowsky, 2011; Milfont, 2012;
Reser et al., 2012b). Australians perceive themselves to be at higher risk
from climate change than New Zealanders and citizens of many other
countries, which may reflect recent experiences of climatic extremes
(Gifford et al., 2009; Agho et al., 2010; Ashworth et al., 2011; Milfont et
al., 2012; Reser et al., 2012c). However, beliefs about climate change
and its risks vary over time, are uneven across society, and reflect media
coverage and bias, political preferences, and gender (ShapeNZ, 2009;
Bacon, 2011; Leviston et al., 2012; Milfont, 2012), which can influence
attitudes to adaptation (Gardner et al., 2010; Gifford, 2011; Reser et al.,
2011; Alexander et al., 2012; Raymond and Spoehr, 2013).
Surveys in Australia between 2007 and 2011 show moderate to high levels
of climate change concern, distress, frustration, resolve, psychological
adaptation, and carbon-reducing behavior (medium evidence, high
agreement; Agho et al., 2010; Reser et al., 2012b,c). About two-thirds of
respondents expected global warming to worsen, with about half very
or extremely concerned that they or their family would be affected
directly. Direct experience with environmental changes or events attributed
to climate change, reported by 45% of respondents, was particularly
influential, but the extent to which resulting distress and concern
translate into support for planned adaptation has not been fully assessed
(Reser et al., 2012a,b).
Box 25-1 (continued)
In addition to raising minimum floor levels and creating coastal setbacks to limit further development in areas at risk, several councils
in Australia and New Zealand have consulted on or attempted to implement managed retreat policies (ECAN, 2005; BSC, 2010; HDC,
2012; KCDC, 2012). These policies remain largely untested in New Zealand, but experience in Australia has shown high litigation
potential and opposing priorities at different levels of government, undermining retreat policies (SCCCWEA, 2009; DCCEE, 2010; Abel
et al., 2011). Mandatory disclosure of information about future risks, community engagement, and policy stability are critical to support
retreat, but existing-use rights, liability concerns, special interests, community resources, place attachment, and divergent priorities at
different levels of government present powerful constraints (high confidence; Hayward, 2008b; Berry and Vella, 2010; McDonald,
2010; Abel et al., 2011; Alexander et al., 2012; Leitch and Robinson, 2012; Macintosh et al., 2013; Reisinger et al., 2013).
1386
Chapter 25 Australasia
25
Frequently Asked Questions
FAQ 25.1 | How can we adapt to climate change
if projected future changes remain uncertain?
Many existing climate change impact assessments in Australia and New Zealand focus on the distant future (2050 to
2
100). When contrasted with more near-term non-climate pressures, the inevitable uncertainty of distant climate impacts
can impede effective adaptation. Emerging best practice in Australasia recognizes this challenge and instead focuses
on those decisions that can and will be made in the near future in any case, along with the “lifetime of those decisions,
a
nd the risk from climate change during that lifetime. Thus, for example, the choice of next year’s annual crop, even
though it is greatly affected by climate, only matters for a year or two and can be adjusted relatively quickly. Even
land-use change among cropping, grazing, and forestry industries has demonstrated significant flexibility in Australasia
over the space of a decade. When the adaptation challenge is reframed as implications for near-term decisions,
uncertainty about the distant future becomes less problematic and adaptation responses can be better integrated
into existing decision-making processes and early warning systems.
(Re)assess
climate-affected
decisions and
overall goals
Potential
impacts within
decision lifetime
Adaptation
options and risk
minimization
Select preferred
option, implement,
and monitor
Decision
Cycle
C
o
m
m
u
n
i
t
y
P
e
r
c
e
p
t
i
o
n
s
&
R
e
a
c
t
i
o
n
S
p
e
c
i
a
l
I
n
t
e
r
e
s
t
s
G
o
v
e
r
n
m
e
n
t
R
e
g
u
l
a
t
i
o
n
s
F
u
n
d
i
n
g
Consequence
Time
Decision Lifetime
Research
Consultations
Revisions
Lead Time
F
r
a
m
i
n
g
,
s
c
o
p
i
n
g
,
a
n
d
r
e
s
o
u
r
c
i
n
g
Decision
Cycle
The Decision-Cycle Context
Adaptation Pathways
Maladaptive Space
Adaptive Space
Maladaptive Space
The Decision Cycle
Figure 25-3 | Adaptation as an iterative risk management process. Individual adaptation decisions comprise well known aspects of risk assessment and management
(top left panel). Each decision occurs within and exerts its own sphere of influence, determined by the lead and consequence time of the decision, and the broader
regulatory and societal influences on the decision (top right panel). A sequence of adaptation decisions creates an adaptation pathway (bottom panel). There is no
single “correct” adaptation pathway, although some decisions, and sequences of decisions, are more likely to result in long-term maladaptive outcomes than others,
but the judgment of outcomes depends strongly on societal values, expectations, and goals.
Continued next page
1387
Australasia Chapter 25
25
Perceived risks and potential losses from climate change depend on values
associated by individuals with specific places, activities, and objects.
Examples from Australia include the value placed on snow cover in the
Snowy Mountains (Gorman-Murray, 2008, 2010), risks to biodiversity
and recreational values in coastal South Australia (Raymond and Brown,
2011), conflicts between human uses and environmental priorities in
national parks (Wyborn, 2009; Roman et al., 2010), and trade-offs between
alternative water supplies and relocation in rural areas (Hurlimann and
Dolnicar, 2011). These and additional studies in Australasia confirm that
the more individuals identify with particular places and their natural
features, the stronger the perceived potential loss but also the greater
the motivation to address environmental threats (e.g., Rogan et al.,
2005; McCleave et al., 2006; Collins and Kearns, 2010; Gosling and
Williams, 2010; Raymond et al., 2011; Russell et al., 2013). This indicates
that ecosystem-based climate change adaptation (see Box CC-EA) can
provide co-benefits for subjective well-being and mental health,
especially for disadvantaged and indigenous communities (Berry et al.,
2010; see also Section 25.8.2).
At the same time, social and cultural values and norms can constrain
adaptation options for communities by limiting the range of acceptable
responses and processes (e.g., place attachment, differing values relating
to near- versus long-term, private versus public, and economic versus
environmental or social costs and benefits, and perceived legitimacy of
institutions). Examples of this are particularly prominent in Australasia
in the coastal zone (e.g., Hayward, 2008a; King et al., 2010; Gorddard
et al., 2012; Hofmeester et al., 2012) and acceptance of water recycling
or pricing (e.g., Pearce et al., 2007; Kouvelis et al., 2010; Mankad and
Tapsuwan, 2011).
Overall, these studies give high confidence that the experience and threat
of climate change and extreme climatic events are having appreciable
psychological impacts, resulting in psychological and subsequent
behavioral adaptations, reflected in high levels of acceptance and
realistic concern; motivational resolve; self-reported changes in thinking,
feeling, and understanding of climate change and its implications; and
behavioral engagement (Reser and Swim, 2011; Reser et al., 2012a,b,c).
However, adequate strategies and systems to monitor trends in
psychological and social impacts, adaptation, and vulnerability are
lacking, and such perspectives remain poorly integrated with and
dominated by biophysical and economic characterizations of climate
change impacts.
25.5. Freshwater Resources
25.5.1. Observed Impacts
Climate change impacts on water represent a cross-cutting issue affecting
people, agriculture, industries, and ecosystems. The challenge of satisfying
multiple demands with a limited resource is exacerbated by the high
interannual and inter-decadal variability of river flows (Chiew and
McMahon, 2002; Peel et al., 2004; Verdon et al., 2004; McKerchar et al.,
2010) particularly in Australia. Declining river flows since the mid-1970s
in far southwestern Australia have led to changed water management
(see Box 11.2 in Hennessy et al., 2007). The unprecedented decline in river
flows during the 1997–2009 “Millennium” drought in southeastern
Australia resulted in low irrigation water allocations, severe water
restrictions in urban centers, suspension of water sharing arrangements,
and major environmental impacts (Chiew and Prosser, 2011; Leblanc et
al., 2012).
25.5.2. Projected Impacts
Figure 25-4 shows estimated changes to mean annual runoff across
Australia for a 1°C global average warming above current levels (Chiew
and Prosser, 2011; Teng et al., 2012). The range of estimates arises mainly
from uncertainty in projected precipitation (Table 25-1). Hydrological
modelling with CMIP3 future climate projections indicates that freshwater
Frequently Asked Questions
FAQ 25.1 (continued)
Some decisions, such as those about long-lived infrastructure and spatial planning and of a public good nature, must
take a long-term view and deal with significant uncertainties and trade-offs between short- and long-term goals
and values. Even then, widely used techniques can help reduce challenges for decision making—including the
“precautionary principle,” “real options,” “adaptive management,“no regrets strategies,or “risk hedging”. These
can be matched to the type of uncertainty but depend on a regulatory framework and institutions that can support
such approaches, including the capacity of practitioners to implement them robustly.
Adaptation is not a one-off action but will take place along an evolving pathway, in which decisions will be revisited
repeatedly as the future unfolds and more information comes to hand (see Figure 25-3). Although this creates learning
opportunities, successive short-term decisions need to be monitored to avoid unwittingly creating an adaptation path
that is not sustainable as climate change continues, or that would cope only with a limited subset of possible climate
futures. This is sometimes referred to as maladaptation. Changing pathways—for example, shifting from ongoing
coastal protection to gradual retreat from the most exposed areas—can be challenging and may require new types
of interactions among governments, industry, and communities.
1388
Chapter 25 Australasia
25
r
esources in far southeastern and far southwest Australia will decline
(high confidence; by 0 to 40% and 20 to 70%, respectively, for 2°C
warming) due to the reduction in winter precipitation (Table 25-1) when
most of the runoff in southern Australia occurs. The percent change in
mean annual precipitation in Australia is generally amplified as a two
to three times larger percent change in mean annual stream flow (Chiew,
2006; Jones et al., 2006).
This can vary, however, with unprecedented declines in flow in far
southeastern Australia in the 1997–2009 drought (Cai and Cowan, 2008;
Potter and Chiew, 2011; Chiew et al., 2013). Higher temperatures and
associated evaporation, tree regrowth following more frequent bushfires
(Kuczera, 1987; Cornish and Vertessy, 2001; Marcar et al., 2006; Lucas
e
t al., 2007), interceptions from farm dams (van Dijk et al., 2006; Lett
et al., 2009), and reduced surface-groundwater connectivity in long dry
spells (Petrone et al., 2010; Hughes et al., 2012) can further accentuate
declines. In the longer-term, water availability will also be affected by
changes in vegetation and surface-atmosphere feedbacks in a warmer
and higher CO
2
environment (Betts et al., 2007; Donohue et al., 2009;
McVicar et al., 2010).
In New Zealand, precipitation changes (Table 25-1) are projected to
lead to increased runoff in the west and south of the South Island
and reduced runoff in the northeast of the South Island, and the east
and north of the North Island (medium confidence). Annual flows of
eastward flowing rivers with headwaters in the Southern Alps (Clutha,
Dry estimate (10th percentile)
Median estimate
Wet estimate (90th percentile)
Change in annual runoff (mm)
5030201055–10–20–30–50
Change in annual runoff (%)
5030201055–10–2030–50
5
0
3
0
2
0
1
0
5
–5
–10
–20
–30
–50
-1-155
-1-188
-1-155
-2-200
-6-6
-3-377
-1-1
00
-5-5
-1-100
-3-3
-2-255
2200
1166
1122
11
00
-1-122
–15
–18
–15
–20
–6
–37
–1
0
–5
–10
–3
–25
20
16
12
1
0
–12
Figure 25-4 | Estimated changes in mean annual runoff for 1°C global average warming above current levels. Maps show changes in annual runoff (percentage change; top
row) and runoff depth (millimeters; bottom row), for dry, median, and wet (10th to 90th percentile) range of estimates, based on hydrological modelling using 15 CMIP3 climate
projections (Chiew et al., 2009; CSIRO, 2009; Petheram et al., 2012; Post et al., 2012). Projections for 2°C global average warming are about twice that shown in the maps (Post
et al., 2011). (Figure adapted from Chiew and Prosser, 2011; Teng et al., 2012).
1389
Australasia Chapter 25
25
W
aimakariri, Rangitata) are projected to increase by 5 to 10 % (median
projection) by 2040 (Bright et al., 2008; Poyck et al., 2011; Zammit and
Woods, 2011) in response to higher alpine precipitation. Most of the
increases occur in winter and spring, as more precipitation falls as rain
and snow melts earlier (Hendrikx et al., 2013). In contrast, the Ashley
River, slightly north of this region, is projected to have little change in
annual flows, with the increase in winter flows offset by reduced summer
flows (Woods et al., 2008). The retreat of glaciers is expected to have
only a minor impact on river flows in the first half of the century (Chinn,
2001; Anderson et al., 2008).
Climate change will affect groundwater through changes in recharge
rates and the relationship between surface waters and aquifers. Dryland
diffuse recharge in most of western, central, and southern Australia is
projected to decrease because of the decline in precipitation, with
increases in the north and some parts of the east because of projected
increase in extreme rainfall intensity (medium confidence; Crosbie et
al., 2010, 2012; McCallum et al., 2010). In New Zealand, a single study
projects groundwater recharge in the Canterbury Plains to decrease
by about 10% by 2040 (Bright et al., 2008). Climate change will also
d
egrade water quality, particularly through increased material washoff
following bushfires and floods (Boxes 25-6, 25-8).
25.5.3. Adaptation
The 1997–2009 drought in southeastern Australia and projected
declines in future water resources in southern Australia are already
stimulating adaptation (Box 25-2). In New Zealand, there is little
evidence of water resources adaptation specifically to climate change.
Water in New Zealand is not as scarce generally and water policy reform
is driven more by pressure to maintain water quality while expanding
agricultural activities, with an increasing focus on collaborative
management (Memon and Skelton, 2007; Memon et al., 2010; Lennox
et al., 2011; Weber et al., 2011) within national guidelines (LWF, 2010;
MfE, 2011). Impacts of climate change on water supply, demand, and
infrastructure have been considered by several New Zealand local
authorities and consultancy reports (Jollands et al., 2007; Williams et
al., 2008; Kouvelis et al., 2010), but no explicit management changes
have yet resulted.
Box 25-2 | Adaptation through Water Resources Policy and Management in Australia
Widespread drought and projections of a drier future in southeastern and far southwest Australia (Bates et al., 2010; CSIRO, 2010;
Potter et al., 2010; Chiew et al., 2011) saw extensive policy and management change in both rural and urban water systems (Hussey
and Dovers, 2007; Bates et al., 2008; Melbourne Water, 2010; DSE, 2011; MDBA, 2011; NWC, 2011; Schofield, 2011). These management
changes provide examples of adaptations, building on previous policy reforms (Botterill and Dovers, 2013). The broad policy framework
is set out in the 2004 National Water Initiative and 2007 Commonwealth Water Act. The establishment of the National Water
Commission (2004) and the Murray-Darling Basin Authority (2008) were major institutional reforms. The National Water Initiative
explicitly recognizes climate change as a constraint on future water allocations. Official assessments (NWC, 2009, 2011) and critiques
(Connell, 2007; Grafton and Hussey, 2007; Byron, 2011; Crase, 2011; Pittock and Finlayson, 2011) have discussed progress and
shortcomings of the initiative, but assessment of its overall success is made difficult by other factors such as ongoing revisions to
allocation plans and time lags to observable impacts.
Rural water reform in southeastern Australia, focused on the Murray-Darling Basin, is currently being implemented. The Murray-
Darling Basin Plan (MDBA, 2011, 2012) will return 2750 GL yr
–1
of consumptive water (about one-fifth of current entitlements) to
riverine ecosystems and develop flexible and adaptive water sharing mechanisms to cope with current and future climates. In 2012,
the Australian government committed more than AU$12 billion nationally to upgrade water infrastructure, improve water use
efficiency, and purchase water entitlements for environmental use. The Basin Plan also includes an environmental watering plan to
optimize environmental outcomes for the Basin. Water markets are a key policy instrument, allowing water use patterns to adapt to
shifting availability and water to move toward higher value uses (NWC, 2010; Kirby et al., 2012). For example, the two-thirds reduction
in irrigation water use over 2000–2009 in the Basin resulted in only 20% reduction in gross agricultural returns, mainly because
water use shifted to more valuable enterprises (Kirby et al., 2012). Elsewhere, catchment management authorities and State agencies
throughout southeastern Australia develop water management strategies to cope with prolonged droughts and climate change (e.g.,
DSE, 2011). Nevertheless, if the extreme dry end of future water projections is realized (Section 25.5.2; Figure 25-4), agriculture and
ecosystems across southeastern and southwestern Australia would be threatened even with comprehensive adaptation (see Sections
25.6.1, 25.7.1-2; Connor et al., 2009; Kirby et al., 2013).
Continued next page
1390
Chapter 25 Australasia
25
25.6. Natural Ecosystems
25.6.1. Inland Freshwater and Terrestrial Ecosystems
Terrestrial and freshwater ecosystems have suffered high rates of habitat
loss and species extinctions since European settlement in both Australia
and New Zealand (Kingsford et al., 2009; Bradshaw et al., 2010; McGlone
et al., 2010; Lundquist et al., 2011; SoE, 2011); many reserves are small
and isolated, and some key ecosystems and species under-represented
(Sattler and Taylor, 2008; MfE, 2010a; SoE, 2011). Many freshwater
ecosystems are pressured from over-allocation and pollution, especially
in southern and eastern coastal regions in Australia (e.g., Ling, 2010).
Additional stresses include erosion, changes in nutrients and fire regimes,
mining, invasive species, grazing, and salinity (Kingsford et al., 2009;
McGlone et al., 2010; SoE, 2011). These increase vulnerability to rapid
climate change and provide challenges for both autonomous and
managed adaptation (Steffen et al., 2009).
25.6.1.1. Observed Impacts
In Australian terrestrial systems, some recently observed changes in the
distribution, genetics, and phenology of individual species, and in the
structure and composition of some ecological communities, can be
attributed to recent climatic trends (medium to high confidence; see
Box 25-3). Uncertainty remains regarding the role of non-climatic drivers,
including changes in atmospheric CO
2
, fire management, grazing, and
land use. The 1997–2009 drought had severe impacts in freshwater
systems in the eastern States and the Murray-Darling Basin (Pittock and
Finlayson, 2011) but, in many freshwater systems, direct climate impacts
are difficult to detect above the strong signal of over-allocation, pollution,
sedimentation, exotic invasions, and natural climate variability (Jenkins
et al., 2011). In New Zealand, few if any impacts on ecosystems have
been directly attributed to climate change rather than variability (Box
25-3; McGlone et al., 2010; McGlone and Walker, 2011). Alpine treelines
in New Zealand have remained roughly stable for several hundred years
(high confidence) despite 0.9°C average warming over the past century
(McGlone and Walker, 2011; Harsch et al., 2012).
25.6.1.2. Projected Impacts
Existing environmental stresses will interact with, and in many cases
be exacerbated by, shifts in mean climatic conditions and associated
change in the frequency or intensity of extreme events, especially fire,
drought, and floods (high confidence; Steffen et al., 2009; Bradstock,
2010; Murphy et al., 2012). Recent drought-related mortality has been
observed for amphibians in southeast Australia (Mac Nally et al., 2009),
savannah trees in northeast Australia (Fensham et al., 2009; Allen et
al., 2010), Mediterranean-type eucalypt forest in southwest Western
Box 25-2 (continued)
Climate change and population growth are the two major factors that influence water planning in Australian capital cities. In
Melbourne, for example, planning has centered on securing new supplies that are more resilient to major climate shocks; increasing
use of alternative sources such as sewage recycling and stormwater for non-potable water; programs to reduce demand; water-
sensitive urban design; and integrated planning that considers climate change impact on water supply, flood risk, and stormwater
and wastewater infrastructures (DSE, 2007; Skinner, 2010; DSE, 2011; Rhodes et al., 2012). Melbourne’s water augmentation program
includes a desalinization plant with a 150 GL yr
–1
capacity (about one-third of the current demand), following the lead of Perth,
where a desalinization plant was established in 2006 because of declining inflows since the mid-1970s (Rhodes et al., 2012).
Melbourne’s water conservation strategies include water efficiency and rebate programs for business and industry, water smart
gardens, dual flush toilets, grey water systems, rainwater tank rebates, free water-efficient showerheads, and voluntary residential
use targets. These conservation measures, together with water use restrictions since the early 2000s, have reduced Melbourne’s total
per capita water use by 40% (Fitzgerald, 2009; Rhodes et al., 2012). Similar programs reduced Brisbane’s per capita water use by
about 50% (Shearer, 2011), while adoption of water recycling and rainwater harvesting resulted in up to 60% water savings in some
parts of Adelaide (Barton and Argue, 2009).
The success of urban water reforms in the face of drought and climate change can be variously interpreted. Increasing supply through
desalinization plants and water reuse schemes reduces the risk of future water shortages and helps cities cope with increasing
population. Uptake of household-scale adaptation options has been locally significant but their long-term sustainability or reversibility
in response to changing drivers and societal attitudes needs further research (Troy, 2008; Brown and Farrelly, 2009; Mankad and
Tapsuwan, 2011). Desalinization plants can be maladaptive because of their energy demand, and the enhancement of mass supply
could create a disincentive for reducing demand or increasing resilience through diversifying supply (Barnett and O'Neill, 2010;
Taptiklis, 2011).
1391
Australasia Chapter 25
25
A
ustralia (Matusik et al., 2013), and eucalypts in sub-alpine regions in
Tasmania (Calder and Kirkpatrick, 2008). Mass die-offs of flying foxes
and cockatoos have been observed during heat waves (Welbergen et
al., 2008; Saunders et al., 2011). These examples provide high confidence
that extreme heat and reduced water availability, either singly or in
combination, will be significant drivers of future population losses and
will increase the risk of local species extinctions in many areas (e.g.,
McKechnie and Wolf, 2010; see also Figure 25-5).
Species distribution modeling (SDM) consistently indicates future range
contractions for Australias native species even assuming optimistic rates
of dispersal, for example, Western Australian Banksia spp. (Fitzpatrick
et al., 2008), koalas (Adams-Hosking et al., 2011), northern macropods
(Ritchie and Bolitho, 2008), native rats (Green, K. et al., 2008), greater
gliders (Kearney et al., 2010b), quokkas (Gibson et al., 2010), platypus
(Klamt et al., 2011), birds (Garnett et al., 2013; van der Wal et al., 2013),
and fish (Bond et al., 2011). In some studies, complete loss of climatically
suitable habitat is projected for some species within a few decades, and
therefore increased risk of local and, perhaps, global extinction (medium
confidence). SDM has limitations (e.g., Elith et al., 2010; McGlone and
Walker, 2011) but is being improved through integration with physiological
(Kearney et al., 2010b) and demographic models (Keith et al., 2008;
Harris et al., 2012), genetic estimates of dispersal capacity (Duckett et
al., 2013), and incorporation into broader risk assessments (e.g., Williams
et al., 2008; Crossman et al., 2012).
In Australia, assessments of ecosystem vulnerability have been based
on observed changes, coupled with projections of future climate in
relation to known biological thresholds and assumptions about adaptive
capacity (e.g., Laurance et al., 2011; Murphy et al., 2012). There is very
high confidence that one of the most vulnerable Australian ecosystems
is the alpine zone because of loss of snow cover, invasions by exotic
species, and changed species interactions (reviewed in Pickering et al.,
2008). There is also high confidence in substantial risks to coastal
wetlands such as Kakadu National Park subject to saline intrusion (BMT
WBM, 2011); tropical savannahs subject to changed fire regimes
(Laurance et al., 2011); inland freshwater and groundwater systems
subject to drought, over-allocation, and altered timing of floods (Pittock
et al., 2008; Jenkins et al., 2011; Pratchett et al., 2011); peat-forming
wetlands along the east coast subject to drying (Keith et al., 2010); and
biodiversity-rich regions such as southwest Western Australia (Yates et
al., 2010a,b) and tropical and subtropical rain forests in Queensland
subject to drying and warming (Stork et al., 2007; Shoo et al., 2011;
Murphy et al., 2012; Hagger et al., 2013).
The very few studies of climate change impacts on biodiversity in New
Zealand suggest that ongoing impacts of invasive species (Box 25-4)
and habitat loss will dominate climate change signals in the short to
medium term (McGlone et al., 2010), but that climate change has the
potential to exacerbate existing stresses (McGlone and Walker, 2011).
There is limited evidence but high agreement that the rich biota of the
alpine zone is at risk through increasing shrubby growth and loss of
herbs, especially if combined with increased establishment of invasive
species (McGlone et al., 2010; McGlone and Walker, 2011). Some cold
water-adapted freshwater fish and invertebrates are vulnerable to
warming (August and Hicks, 2008; Winterbourn et al., 2008; Hitchings,
2009; McGlone and Walker, 2011) and increased spring flooding may
i
ncrease risks for braided-river bird species (MfE, 2008b). For some
restricted native species, suitable habitat may increase with warming
(e.g., native frogs; Fouquet et al., 2010) although limited dispersal ability
will limit range expansion. Tuatara populations are at risk as warming
increases the ratio of males to females (Mitchell et al., 2010), although
the lineage has persisted during higher temperatures in the geological
past (McGlone and Walker, 2011).
25.6.1.3. Adaptation
High levels of endemism in both countries (Lindenmayer, 2007; Lundquist
et al., 2011) are associated with narrow geographic ranges and associated
climatic vulnerability, although there is greater scope for adaptive
dispersal to higher elevations in New Zealand than in Australia.
Anticipated rates of climate change, together with fragmentation of
remaining habitat and limited migration options in many regions
(Steffen et al., 2009; Morrongiello et al., 2011), will limit in situ adaptive
capacity and distributional shifts to more climatically suitable areas for
many species (high confidence). Significant local and global losses of
species, functional diversity, and ecosystem services, and large-scale
changes in ecological communities, are anticipated (e.g., Dunlop et al.,
2012; Gallagher et al., 2012b; Murphy et al., 2012).
There is increasing recognition in Australia that rapid climate change
has fundamental implications for traditional conservation objectives
(e.g., Steffen et al., 2009; Prober and Dunlop, 2011; Dunlop et al., 2012;
Murphy et al., 2012). Research on impacts and adaptation in terrestrial
and freshwater systems has been guided by the National Adaptation
Research Plans (Hughes et al., 2010; Bates et al., 2011) and by research
undertaken within the CSIRO Climate Adaptation Flagship. Climate
change adaptation plans developed by many levels of government and
Natural Resource Management (NRM) bodies, supported by substantial
Australian government funding, have identified priorities that include
identification and protection of climatic refugia (Davis et al., 2013;
Reside et al., 2013); restoration of riparian zones to reduce stream
temperatures (Davies, 2010; Jenkins et al., 2011); construction of levees
to protect wetlands from saltwater intrusion (Jenkins et al., 2011);
reduction of non-climatic threats such as invasive species to increase
ecosystem resilience (Kingsford et al., 2009); ecologically appropriate
fire regimes (Driscoll et al., 2010); restoration of environmental flows
in major rivers (Kingsford and Watson, 2011; Pittock and Finlayson,
2011); protecting and restoring habitat connectivity in association with
expansion of the protected area network (Dunlop and Brown, 2008;
Mackey et al., 2008; Taylor and Philp, 2010; Prowse and Brook, 2011;
Maggini et al., 2013); and active interventionist strategies such as
assisted colonization to reduce probability of species extinctions
(Burbidge et al., 2011; McIntyre, 2011) or restore ecosystem services
(Lunt et al., 2013). Few specific measures have been implemented and
thus their effectiveness cannot yet be assessed. Biodiversity research
and management in New Zealand to date has taken little account of
climate change-related pressures and continues to focus largely on
managing pressures from invasive species and predators, freshwater
pollution, exotic diseases, and halting the decline in native vegetation,
although a number of specific recommendations have been made to
improve ecosystem resilience to future climate threats (McGlone et al.,
2010; McGlone and Walker, 2011).
1392
Chapter 25 Australasia
25
C
limate change responses in other sectors may have beneficial as well
as adverse impacts on biodiversity, but few tools to assess risks from
an integrated perspective have been developed (Section 25.9.1; Box
25-10). Assessments of the impacts of climate change on the provision
of ecosystem services (such as pollination and erosion control) via
impacts on terrestrial and freshwater ecosystems are generally lacking.
Similarly, the concept of Ecosystem-based Adaptation—the role of
healthy, well-functioning ecosystems in increasing the resilience of
human sectors to the impacts of climate change (see Chapters 4 and 5;
Box CC-EA)—is relatively unexplored.
25.6.2. Coastal and Ocean Ecosystems
Australia’s 60,000 km coastline spans tropical waters in the north to
cool temperate waters off Tasmania and the sub-Antarctic islands with
sovereign rights over approximately 8.1 million km
2
, excluding the
Australian Antarctic Territory (Richardson and Poloczanska, 2009). New
Zealand has approximately 18,000 km of coastline, spanning subtropical
to sub-Antarctic waters, and the world’s fifth largest Exclusive Economic
Zone at 4.2 million km
2
(Gordon et al., 2010). The marine ecosystems
of both countries are considered hotspots of global marine biodiversity
with many rare, endemic, and commercially important species (Hoegh-
Guldberg et al., 2007; Blanchette et al., 2009; Gordon et al., 2010;
Gillanders et al., 2011; Lundquist et al., 2011). The increasing density of
coastal populations (see Section 25.3) and stressors such as pollution
and sedimentation from settlements and agriculture will intensify non-
climate stressors in coastal areas (high confidence; e.g., Russell et al.,
2009). Coastal habitats provide many ecosystem services including
coastal protection (Arkema et al., 2013) and carbon storage, particularly
in seagrass, saltmarsh, and mangroves, which could become increasingly
important for mitigation (e.g., Irving et al., 2011). Coastal ecosystems
occupy less than 1% of the land mass but may account for 39% of
Australia’s average national annual carbon burial (estimated total: 466
millions tonnes CO
2
-eq per year; Lawrence et al., 2012).
25.6.2.1. Observed Impacts
There is high confidence that climate change is already affecting the
oceans around Australia (Pearce and Feng, 2007; Poloczanska et al., 2007;
Lough and Hobday, 2011) and warming the Tasman sea in northern New
Zealand (Sutton et al., 2005; Lundquist et al., 2011); average climate
zones have shifted south by more than 200 km along the northeast and
about 100 km along the northwest Australian coasts since 1950 (Lough,
2008). The rate of warming is even faster in southeast Australia, with
a poleward advance of the East Australia Current of approximately
350 km over the past 60 years (Ridgway, 2007). Based on elevated rates
of ocean warming, southwest and southeast Australia are recognized
as global warming hotspots (Wernberg et al., 2011). It is virtually
certain that the increased storage of carbon by the ocean will increase
acidification in the future, continuing the observed trends of the past
decades in Australia as elsewhere (Howard et al., 2012; see also WGI
AR5 Sections 3.8, 6.44).
Recently observed changes in marine systems around Australia are
consistent with warming oceans (high confidence; Box 25-3). Examples
i
nclude changes in phytoplankton productivity (Thompson et al., 2009;
Johnson et al., 2011); species abundance of macroalgae (Johnson et al.,
2011); growth rates of abalone (Johnson et al., 2011), southern rock
lobster (Pecl et al., 2009; Johnson et al., 2011), coastal fish (Neuheimer
et al., 2011), and coral (De’ath et al., 2009); life cycles of southern rock
lobster (Pecl et al., 2009) and seabirds (Cullen et al., 2009; Chambers
et al., 2011); and distribution of subtidal seaweeds (Johnson et al., 2011;
Wernberg et al., 2011; Smale and Wernberg, 2013), plankton (Mcleod
et al., 2012), fish (Figueira et al., 2009; Figueira and Booth, 2010; Last
et al., 2011; Madin et al., 2012), sea urchins (Ling et al., 2009), and
intertidal invertebrates (Pitt et al., 2010).
Habitat-related impacts are more prevalent in northern Australia
(Pratchett et al., 2011), while distribution changes are reported more
often in southern waters (Madin et al., 2012), particularly southeast
Australia, where warming has been greatest. The 2011 marine heat
wave in Western Australia caused the first-ever reported bleaching at
Ningaloo reef (Abdo et al., 2012; Feng et al., 2013), resulting in coral
mortality (Moore et al., 2012; Depczynski et al., 2013) and changes in
community structure and composition (Smale and Wernberg, 2013;
Wernberg et al., 2013). About 10% of the observed 50% decline in coral
cover on the Great Barrier Reef since 1985 has been attributed to
bleaching, the remainder to cyclones and predators (De’ath et al., 2012).
Changes in distribution and abundance of marine species in New
Zealand are primarily linked to ENSO-related variability that dominates
in many time series (Clucas, 2011; Lundquist et al., 2011; McGlone and
Walker, 2011; Schiel, 2011), although water temperature is also important
(e.g., Beentjes and Renwick, 2001). New Zealand fisheries exported
more than NZ$1.5 billion worth of product in 2012 (SNZ, 2013) and
variability in ocean circulation and temperature plays an important role
in local fish abundance (e.g., Chiswell and Booth, 2005; Dunn et al.,
2009); no climate change impacts have been reported at this stage
(Dunn et al., 2009), although this may be due to insufficient monitoring.
25.6.2.2. Projected Impacts
Even though evidence of climate impacts on coastal habitats is limited
to date, confidence is high that negative impacts will arise with continued
climate change (Lovelock et al., 2009; McGlone and Walker, 2011; Traill
et al., 2011; Chapter 6). Some coastal habitats such as mangroves are
projected to expand further landward, driven by sea level rise and
exacerbated by soil subsidence if rainfall declines (medium confidence;
Traill et al., 2011), although this may be at the expense of saltmarsh
and constrained in many regions by the built environment (DCC, 2009;
Lovelock et al., 2009; Rogers et al., 2012). Estuarine habitats will be
affected by changing rainfall or sediment discharges, as well as
connectivity to the ocean (high confidence; Gillanders et al., 2011). Loss
of coastal habitats and declines in iconic species will result in substantial
impacts on coastal settlements and infrastructure from direct impacts
such as storm surge, and will affect tourism (medium confidence; Section
25.7.5).
Changes in temperature and rainfall, and sea level rise, are expected to
lead to secondary effects, including erosion, landslips, and flooding,
affecting coastal habitats and their dependent species, for example, loss
1393
Australasia Chapter 25
25
o
f habitat for nesting birds (high confidence; Chambers et al., 2011).
Increasing ocean acidification is expected to affect many taxa (medium
confidence; see also Box CC-OA; Chapters 6, 30) including corals
(Fabricius et al., 2011), coralline algae (Anthony et al., 2008), calcareous
plankton (Richardson et al., 2009; Thompson et al., 2009; Hallegraeff,
2010), reef fishes (Munday et al., 2009; Nilsson et al., 2012), bryozoans,
and other benthic calcifiers (Fabricius et al., 2011). Deep-sea scleractinian
corals are also expected to decline with ocean acidification (Miller et
al., 2011).
The AR4 identified the Great Barrier Reef as highly vulnerable to both
warming and acidification (Hennessy et al., 2007). Recent observations
of bleaching (GBRMPA, 2009a) and reduced calcification in both the
Great Barrier Reef and other reef systems (Cooper et al., 2008; De’ath
et al., 2009; Cooper et al., 2012), along with model and experimental
studies (Hoegh-Guldberg et al., 2007; Anthony et al., 2008; Veron et al.,
2009) confirm this vulnerability (see also Box CC-CR). The combined
impacts of warming and acidification associated with atmospheric CO
2
concentrations in excess of 450 to 500 ppm are projected to be associated
with increased frequency and severity of coral bleaching, disease incidence
and mortality, in turn leading to changes in community composition
and structure including increasing dominance by macroalgae (high
confidence; Hoegh-Guldberg et al., 2007; Veron et al., 2009). Other
stresses, including rising sea levels, increased cyclone intensity, and
nutrient-enriched and freshwater runoff, will exacerbate these impacts
(high confidence; Hoegh-Guldberg et al., 2007; Veron et al., 2009;
GBRMPA, 2013). Thermal thresholds and the ability to recover from
bleaching events vary geographically and between species (e.g., Diaz-
Pulido et al., 2009) but evidence of the ability of corals to adapt to rising
temperatures and acidification is limited and appears insufficient to
offset the detrimental effects of warming and acidification (robust
evidence, medium agreement; Hoegh-Guldberg, 2012; Howells et al.,
2013; Box CC-CR).
Under all SRES scenarios and a range of CMIP3 models, pelagic fishes
such as sharks, tuna, and billfish are projected to move further south
on the east and west coasts of Australia (high confidence; Hobday,
2010). These changes depend on sensitivity to water temperature, and
may lead to shifts in species-overlap with implications for by-catch
management (Hartog et al., 2011). Poleward movements are also
projected for coastal fish species in Western Australia (Cheung et al.,
2012) and a complex suite of impacts are expected for marine mammals
(Schumann et al., 2013). A strengthening East Auckland Current in
northern New Zealand is expected to promote establishment of tropical
or subtropical species that currently occur as vagrants in warm La Niña
years (Willis et al., 2007). Such shifts suggest potentially substantial
changes in production and profit of both wild fisheries (Norman-Lopez
et al., 2011) and aquaculture species such as salmon, mussels, and oysters
(medium confidence; Hobday et al., 2008; Hobday and Poloczanska,
2010). Ecosystem models also project changes to habitat and fisheries
production (low confidence; Fulton, 2011; Watson et al., 2012).
25.6.2.3. Adaptation
In Australia, research on marine impacts and adaptation has been
guided by the National Adaptation Research Plan for Marine Biodiversity
a
nd Resources (Mapstone et al., 2010), programs within the CSIRO
Climate Adaptation Flagship, and the Great Barrier Reef Marine Park
Authority (GBRMPA, 2007). Limits to autonomous adaptation are
unknown for almost all species, although limited experiments suggests
capacity for response on a scale comparable to projected warming for
some species (e.g., coral reef fish; Miller et al., 2012) and not others
(e.g., Antarctic krill; Kawaguchi et al., 2013). Planned adaptation options
include removal of human barriers to landward migration of species,
beach nourishment, management of environmental flows to maintain
estuaries (Jenkins et al., 2010), habitat provision (Hobday and Poloczanska,
2010), assisted colonization of seagrass and species such as turtles (e.g.,
Fuentes et al., 2009), and burrow modification for nesting seabirds
(Chambers et al., 2011).
For southern species on the continental shelf, options are more limited
because suitable habitat will not be present—the next shallow water
to the south is Macquarie Island. There is low confidence about the
adequacy of autonomous rates of adaptation by species, although recent
experiments with coral reef fish suggest that some species may adapt
to the projected climate changes (Miller et al., 2012).
Management actions to increase coral reef resilience include reducing
fishing pressure on herbivorous fish, protecting top predators, managing
runoff quality, and minimizing other human disturbances, especially
through marine protected areas (Hughes et al., 2007; Veron et al., 2009;
Wooldridge et al., 2012). Such actions will slow, but not prevent, long-
term degradation of reef systems once critical thresholds of ocean
temperature and acidity are exceeded (high confidence), and so novel
options, including assisted colonization and shading critical reefs, have
been proposed but remain untested at scale (Rau et al., 2012). Seasonal
forecasting can also prepare managers for bleaching events (Spillman,
2011).
Adaptation by the fishing industry to shifting distributions of target
species is considered possible by most stakeholders (e.g., southern rock
lobster fishery; Pecl et al., 2009). Assisted colonization to maintain
production in the face of declining recruitment may also be possible for
some high value species, and has been trialed for the southern rock
lobster (Green, B.S. et al., 2010). Options for aquaculture include disease
management, alternative site selection, and selective breeding (Battaglene
et al., 2008), but implementation is only preliminary. Marine protected
area planning is not explicitly considering climate change in either country,
but reserve performance will be affected by projected environment
shifts and novel combinations of species, habitats, and human pressures
(Hobday, 2011).
25.7. Major Industries
25.7.1. Production Forestry
Australia has about 149 Mha of forests, including woodlands. Two Mha
are plantations and 9.4 Mha multiple-use native forests, and forestry
contributes around AU$7 billion annually to GDP (ABARES, 2012). New
Zealand’s plantation estate in production forests comprises about 1.7
Mha (90% Pinus radiata), with recent contractions due to increased
profitability of dairying (FOA and MPI, 2012; MfE, 2013).
1394
Chapter 25 Australasia
25
Box 25-3 | Impacts of a Changing Climate in Natural and Managed Ecosystems
Observed changes in species, and in natural and managed ecosystems (Sections 25.6.1-2, 25.7.2) provide multiple lines of evidence
of the impacts of a changing climate. Examples of observations published since the AR4 are shown in Table 25-3.
Continued next page
Type of change
and nature of
evidence
Examples
Time scale of
observations
Confi dence in
the detection
of biological
change
Potential climate change
driver(s)
Confi dence
in the role
of climate vs
other drivers
Morphology
L
imited evidence
(
one study)
Declining body size of southeast Australian
passerine birds, equivalent to ~7° latitudinal shift
(Gardner et al., 2009)
About 100 years Medium: Trend
signifi cant
for 4 out of 8
species; 2 other
species show
same trend but
n
ot statistically
s
ignifi cant
Warming air temperatures about
1.0°C over same period
Medium:
Nutritional cause
discounted
Geographic distribution
High agreement,
robust evidence for
many marine species
and mobile terrestrial
species
Southerly range extension of the barrens-forming
s
ea urchin Centrostephanus rodgersii from the New
South Wales coast to Tasmania; ow on impacts to
m
arine communities including lobster shery; shift
of 160 km per decade over 30 years (Ling, 2008;
Ling et al., 2008, 2009; Banks et al., 2010)
About 30 50 years
(
rst recorded in
Tasmania in late 1970s)
High Increased SST, ocean warming
i
n southeast Australia, increased
southerly penetration of the
E
AC, 350 km over 60 years
High
45 fi sh species, representing 27 families (about
30% of the inshore fi sh families occurring in the
region), exhibited major distributional shifts in
Tasmania (Last et al., 2011).
Distributions from
late 1880s, 1980s and
present (1995 now)
High Increased SST in southeast
Australia, increased southerly
penetration of the EAC
Medium:
Changed fi shing
practices have
potentially
contributed to
trends.
Southward range shift of intertidal species
(average minimum distance 116 km) off west
coast of Tasmania; 55% species recorded at more
southerly sites; only 3% species expanded to
more northerly sites (Pitt et al., 2010).
About 50 years; sites
resampled 2007– 2008,
compared with 1950s
Medium
Increased SST in southeast
Australia (average 0.22°C per
decade), increased southerly
penetration of the EAC, 350 km
over 60 years
Medium
Life cycles
Robust evidence,
medium agreement;
increasing
documentation of
advances in phenology
in some species
(mainly migration
and reproduction in
birds, emergence in
butterfl ies, owering
in plants) but also
signifi cant trends
toward later life cycle
events in some taxa;
see meta-analysis for
Southern Hemisphere
phenology (Chambers
et al., 2013a)
Signifi cant advance in mean emergence date of
1.5 days per decade (1941– 2005) in the Common
Brown Butterfl y Heteronympha merope in
Australia (Kearney et al., 2010)
65 years High Increase in local air
temperatures of 0.16°C per
decade (1945 2007)
High: Advance
consistent with
physiologically
based model
of temperature
infl uence on
development
Advances in spring phenology of migratory birds,
and both advances and delays in phenology
in other seasons at multiple Australian sites:
meta-analysis of 52 species and 145 data sets
(Chambers et al., 2013b)
Multiple time periods
from 1960s, all
included 1990s and
2000s
High Local climate trends (increasing
air temperature, decreased rain
days) were more important
than broad-scale drivers such as
the Southern Oscillation Index.
Strongest associations were with
decreased rain days.
High: No other
potential
confounding
factors identifi ed
Earlier wine-grape ripening at 9 of 10 sites in
Australia (Webb, L. B. et al., 2012)
Multiple time periods
up to 64 years (average
41 years)
High Increased length of growing
season, increased average
temperature, and reduced soil
moisture
Medium:
Changed
husbandry
techniques,
resulting in lower
crop yields, may
have contributed
to trend.
Timing of migration of glass eels, Anguilla spp.,
advanced by several weeks in Waikato River,
North Island, New Zealand (Jellyman et al., 2009).
30 years (2004 2005
compared to 1970s)
Medium Warming water temperatures in
spawning grounds
Low: Changes
in discharge
discounted as
contributing
factor
Table 25-3 | Examples of detected changes in species, natural and managed ecosystems, consistent with a climate change signal, published since the AR4. Confi dence
in detection of change is based on the length of study and the type, amount, and quality of data in relation to the natural variability in the particular species or system.
Confi dence in the role of climate being a major driver of the change is based on the extent to which the detected change is consistent with that expected under climate
change, and to which other confounding or interacting non-climate factors have been considered and been found insuffi cient to explain the observed change. (SST =
sea surface temperature; EAC = East Australian Current.)
1395
Australasia Chapter 25
25
25.7.1.1. Observed and Projected Impacts
Existing climate variability and other confounding factors have so far
prevented the detection of climate change impacts on forests. Modeled
projections are based on ecophysiological responses of forests to CO
2
,
water, and temperatures. In Australia, potential changes in water
availability will be most important (very high confidence; e.g., reviews
by Battaglia et al., 2009; Medlyn et al., 2011b). Modeling future
distributions or growth rates indicate that plantations in southwest
Western Australia are most at risk due to declining rainfall, and there is
high confidence that plantation growth will be reduced by temperature
increases in hotter regions, especially where species are grown at the
upper range of their temperature tolerances (Medlyn et al., 2011a).
Moderate reductions in rainfall and increased temperature could be
offset by fertilization from increasing CO
2
(limited evidence, medium
agreement; Simioni et al., 2009). In cool regions where water is not
limiting, higher temperatures could benefit production (Battaglia et al.,
2009). In New Zealand, temperatures are mostly sub-optimal for growth
of P. radiata and water relations are generally less limiting (Kirschbaum
and Watt, 2011). Warming is expected to increase P. radiata growth in
the cooler south (very high confidence), whereas in the warmer north,
temperature increases can reduce productivity, but CO
2
fertilization may
offset this (medium confidence; Kirschbaum et al., 2012).
Modeling studies are limited by their reliance on key assumptions which
are difficult to verify experimentally, for example, the degree to which
Box 25-3 (continued)
Type of change
and nature of
evidence
Examples
Time scale of
observations
Confi dence in
the detection
of biological
change
Potential climate change
driver(s)
Confi dence
in the role
of climate vs
other drivers
Marine productivity
L
imited evidence,
medium agreement
Otolith (“ear stone”) analyses in long-lived Pacifi c
sh indicates signifi cantly increased growth
r
ates for shallow-water species (< 250m) (3 of
3
species), reduced growth rates of deep-water
(
>1000m) species (3 of 3 species); no change
o
bserved in the 2 intermediate-depth species
(
Thresher et al., 2007).
Birth years ranged
1
861–1993 (fi sh 2–128
y
ears old)
High Increasing growth rates in
s
pecies in top 250 m associated
w
ith warming SST, declining
g
rowth rates in species >1000
m
associated with long-term
c
ooling (as indicated by Mg / Ca
r
atios and change in
18
O
in deep
water corals)
Medium:
C
hanged fi shing
p
ressure may
h
ave contributed
t
o trend.
A
bout a 50% decline in growth rate and biomass
o
f spring phytoplankton bloom in western Tasman
Sea (Thompson et al., 2009)
6
0-year data set;
d
ecline recorded over
period 1997– 2007
H
igh Increased SST and extension of
t
he EAC associated with reduced
nutrient availability
M
edium
V
egetation change
Limited agreement and
e
vidence; interacting
impacts of changed
land practices; altered
re regimes, increasing
a
tmospheric CO
2
c
oncentration and
climate trends diffi cult
to disentangle
E
xpansion of monsoon rainforest at expense of
eucalypt savannah and grassland in Northern
T
erritory, Australia (Banfai and Bowman, 2007;
Bowman et al., 2010)
A
bout 40 years Medium Increases in rainfall and
atmospheric CO
2
M
edium:
Changes in fi re
r
egimes and land
management
p
ractices may
have contributed
to trend.
Net increase in mire wetland extent (10.2%) and
corresponding contraction of adjacent eucalypt
woodland in seven sub-catchments in southeast
Australia (Keith et al., 2010)
Weather data covers
> 40 years (depending
on parameter);
vegetation mapping
from 1961 to 1998.
Medium Decline in evapo-transpiration Low: Resource
exploitation,
re history, and
autogenic mire
development
discounted
Freshwater
communities
Limited evidence
(one study)
Decline in families of macroinvertebrates that
favor cooler, faster-fl owing habitats in New South
Wales streams and increase in families favoring
warmer and more lentic conditions (Chessman,
2009)
13 years (1994 2007) Medium Increasing water temperatures
and declining fl ows
Low: Variation
in sampling,
changes in water
quality, impacts
of impoundment
and water
extraction may
have contributed
to trends.
Disease
Limited evidence,
robust agreement
Emergence and increased incidence of coral
diseases including white syndrome (since 1998)
and black band disease (since 1993 1994) (Bruno
et al., 2007; Sato et al., 2009; Dalton et al., 2010)
1998 onwards Medium Increasing SST High
Coral reefs
Robust evidence,
high agreement
Multiple mass bleaching events since 1979 (see
Sections 25.6.2 and 30.5)
1979 onwards High Increasing SST High
Calcifi cation of Porites on GBR declined 21%
(1971– 2003, 4 reefs; Cooper et al., 2008); about
11% (1990 2005, 69 reefs; De’ath et al., 2009)
1 9 7 1 2 0 0 3 ;
1990 2005
High Increasing SST High: Changes
in water quality
discounted
Table 25-3 (continued)
1396
Chapter 25 Australasia
25
p
hotosynthesis remains stimulated under elevated CO
2
(
Battaglia et al.,
2009). Most studies also exclude impacts of pests, diseases, weeds, fire,
and wind damage that may change adversely with climate. Fire, for
instance, poses a significant threat in Australia and is expected to worsen
with climate change (see Box 25-6), especially for the commercial forestry
plantations in the southern winter-rainfall regions (Williams et al., 2009;
Clarke et al., 2011). In New Zealand, changes in biotic factors are
particularly important as they already affect plantation productivity.
Dothistroma blight, for instance, is a serious pine disease with a
temperature optimum that coincides with New Zealand’s warmer, but
not warmest, pine-growing regions; under climate change, its severity
is, therefore, expected to reduce in the warm central North Island but
increase in the cooler South Island (high confidence) where it could offset
temperature-driven improved plantation growth (Watt et al., 2011a). There
is medium evidence and high agreement of similar future southward shifts
in the distribution of existing plantation weed, insect pest, and disease
species in Australia (see review in Medlyn et al., 2011b).
25.7.1.2. Adaptation
Depending on the extent of climate changes and plant responses to
increasing CO
2
, the above studies provide limited evidence but high
agreement of potential net increased productivity in many areas, but
only where soil nutrients are not limiting. Adaptation strategies include
changes to species or provenance selection toward trees better adapted
to warmer conditions, or adopting different silvicultural options to increase
resilience to climatic or biotic stresses, such as pest challenges (White et
al., 2009; Booth et al., 2010; Singh et al., 2010; Wilson and Turton, 2011a).
The greatest barriers to long-term adaptation planning are incomplete
knowledge of plant responses to increased CO
2
and uncertainty in
regional climate scenarios (medium evidence, high agreement; Medlyn
et al., 2011b). The rotation time of plantation forests of about 30 years or
more makes proactive adaptation important but also challenging.
25.7.2. Agriculture
Australia produces 93% of its domestic food requirements and exports
76% of agricultural production (PMSEIC, 2010a). New Zealand agriculture
contributes about 56% of total export value and dairy products 27%;
95% of dairy products are exported (SNZ, 2012b). Agricultural production
is sensitive to climate (especially drought; Box 25-5) but also to many
non-climate factors such as management, which thus far has limited
both detection and attribution of climate-related changes (see Chapters
7, 18; Webb, L.B. et al., 2012; Darbyshire et al., 2013). Because the region
is a major exporter—providing, for example, more than 40% of the
world trade in dairy products—changes in production conditions in the
region have a major influence on world supply (OECD, 2011). This
implies that climate change impacts could have consequences for food
security not just locally but even globally (Qureshi et al., 2013a).
25.7.2.1. Projected Impacts and Adaptation—Livestock Systems
Livestock grazing dominates land use by area in the region. At the
Australian national level, the net effect of a 3°C temperature increase
(
from a 1980–1999 baseline) is expected to be a 4% reduction in gross
value of the beef, sheep, and wool sector (McKeon et al., 2008). Dairy
productivity is projected to decline in all regions of Australia other than
Tasmania under a mid-range (A1B) climate scenario by 2050 (Hanslow
et al., 2013). Projected changes in national pasture production for
dairy, sheep, and beef pastures in New Zealand range from an average
reduction of 4% across climate scenarios for the 2030s (Wratt et al.,
2008) to increases of up to 4% for two scenarios in the 2050s (Baisden
et al., 2010) when the models included CO
2
fertilization and nitrogen
feedbacks.
Studies modeling seasonal changes in fodder supply show greater
sensitivity in animal production to climate change and elevated CO
2
than models using annual average production, with some impacts
expected even under modest warming (high confidence) in both New
Zealand (Lieffering et al., 2012) and Australia (Moore and Ghahramani,
2013). Across 25 sites in southern Australia (an area that produces 85%
of sheep and 40% of beef production by value) modeled profitability
declined at most sites by the 2050s because of a shorter growing season
due to changes in both rainfall and temperature (Moore and Ghahramani,
2013). In New Zealand, projected changes in seasonal pasture growth
drove changes in animal production at four sites representing the main
areas of sheep production (Lieffering et al., 2012). In Hawke’s Bay,
changes in stock number and the timing of grazing were able to maintain
farm income for a period in the face of variable forage supply but not
in the longer term. In Southland and Waikato, projected increases in
early spring pasture growth posed management problems in maintaining
pasture quality, yet, if these were met, animal production could be
maintained or increased. The temperature-humidity index (THI), an
indicator of potential heat stress for animals, increased from 1960 to
2008 in the Murray Dairy region of Australia and further increases and
reductions in milk production are projected (Nidumolu et al., 2011).
Shading can substantially reduce, but not avoid, the temperature and
humidity effects that produce a high THI (Nidumolu et al., 2011).
Rainfall is a key determinant of interannual variability in production and
profitability of pastures and rangelands (Radcliffe and Baars, 1987;
Steffen et al., 2011) yet remains the most uncertain change. In northern
Australia, incremental adaptation may be adequate to manage risks
of climate change to the grazing industry but an increasing frequency
of droughts and reduced summer rainfall will potentially drive the
requirement for transformational change (Cobon et al., 2009). Rangelands
that are currently water-limited are expected to show greater sensitivity
to temperature and rainfall changes than nitrogen-limited ones
(Webb, N.P. et al., 2012). The “water-sparing” effect of elevated CO
2
(offsetting reduced water availability from reduced rainfall and
increased temperatures) is invoked in many impact studies but does not
always translate into production benefits (Kamman et al., 2005; Newton
et al., 2006; Stokes and Ash, 2007; Wan et al., 2007). The impact of
elevated CO
2
on forage production, quality, nutrient cycling, and water
availability remains the major uncertainty in modeling system responses
(McKeon et al., 2009; Finger et al., 2010); recent findings of grazing
impacts on plant species composition (Newton et al., 2013) and nitrogen
fixation (Watanabe et al., 2013) under elevated CO
2
have added to this
uncertainty. New Zealand agro-ecosystems are subject to erosion
processes strongly driven by climate; greater certainty in projections of
rainfall, particularly storm frequency, are needed to better understand
1397
Australasia Chapter 25
25
climate change impacts on erosion and consequent changes in the
ecosystem services provided by soils (Basher et al., 2012).
25.7.2.2. Projected Impacts and Adaptation—Cropping
Experiments with elevated CO
2
at two sites with different temperatures
have shown a wide range in the response of current wheat cultivars
(Fitzgerald et al., 2010). Modeling suggests there is the potential to
increase New Zealand wheat yields under climate change with
appropriate choices of cultivars and sowing dates (high confidence;
Teixeira et al., 2012). In Australia, the selection of appropriate cultivars
and sowing times is projected to result in increased wheat yields in high
rainfall areas such as southern Victoria under climate change and in
maintenance of current yields in some areas expected to be drier (e.g.,
northwestern Victoria; O’Leary et al., 2010). However, if extreme low
rainfall scenarios are realized in areas such as South Australia then
changes in cultivars and fertilizer applications are not expected to
maintain current yields by 2080 (Luo et al., 2009). Under the more
severe climate scenarios and without adaptation, Australia could become
a net importer of wheat (Howden et al., 2010). One caveat to modeling
studies is that an intercomparison of 27 wheat models found large
differences between model outputs for already dry and hot Australian
sites in response to increasing CO
2
and temperature (Asseng et al., 2013;
Carter, 2013).
Rice production in Australia is largely dependent on irrigation, and
climate change impacts will strongly depend on water availability and
price (Gaydon et al., 2010). Sugarcane is also strongly water dependent
(Carr and Knox, 2011); yields may increase where rainfall is unchanged
or increased, but rising temperatures could drive up evapotranspiration
and increase water use (medium confidence; Park et al., 2010).
Box 25-4 | Biosecurity
Biosecurity is a high priority for Australia and New Zealand given the economic importance of biologically based industries and risks
to endemic species and iconic ecosystems. The biology and potential risk from invasive and native pathogenic species will be altered
by climate change (high confidence; Roura-Pascual et al., 2011), but impacts may be positive or negative depending on the particular
system.
Consequence Projected change Organism /ecosystem affected
Altered mechanisms of
transport and introduction
Increased risk of introduction of Asiatic citrus psyllid ( Diaphorina citri), vector of the disease
huanglongbing
1
Australian citrus industry and native citrus
and other rutaceous species and endemic
psyllid fauna
Altered distribution of
existing invasive and
pathogenic species
N
assella neesiana (Chilean needle grass): Increased droughts favor establishment.
2
M
anaged pasture in New Zealand
W
arming and drying may encourage the spread of existing invasives such as Pheidole
m
egacephala in New Zealand and provide suitable conditions for other exotic ant species if they
i
nvade.
3
H
uman health and potentially agricultural
a
nd natural ecosystems
Reduced climatic suitability for exotic invasive grasses in Australia (11 species including Nassella
s
p.)
4
Australian rangeland
R
ange of the invasive weed Lantana camara (lantana) projected to extend from north Australia to
Victoria, southern Australia, and Tasmania
5
M
ultiple
P
rojected increases in the range of three recently naturalized subtropical plants ( Archontophoenix
cunninghamiana, Psidium guajava, Scheffl era actinophylla)
6
N
ative ecosystems in New Zealand
Altered climatic
constraints on invasive
and pathogenic species
Queensland fruit fl y ( Bactrocera tryoni) moving southwards
7
Australian horticulture
Signifi cant association between amphibian declines in upland rainforests of north Queensland and
three consecutive years of warm weather suggests future warming could increase the vulnerability
of frogs to chytridiomycosis caused by the chytrid fungus Batrachochytrium dendrobatadis.
8
Native frogs
Altered impact of existing
invasive and pathogenic
species
Fusarium pseudograminearum causing crown rot increases under elevated CO
2
.
9
Australian wheat
Increased abundance of the root-feeding nematode Longidorus elongatus under elevated CO
2
10
New Zealand pasture
Increased severity of Swiss needle cast disease caused by Phaeocryptopus gaeumannii
11
Douglas fi r plantations in New Zealand,
impact more severe in North Island
Altered effectiveness of
management strategies
Light brown apple moth, Epiphyas postvittana (Walker) ( Lepidoptera: Tortricidae) reduction in
natural enemies due to asynchrony and loss of host species
12
Australian horticulture
Projected changes in the effi cacy of fi ve biological control systems demonstrating a range of
potential disruption mechanisms
13
Pastoral and horticultural systems in New
Zealand
Table 25-4 | Examples of potential consequences of climate change for invasive and pathogenic species relevant to Australia and New Zealand, with consequence
categories based on Hellman et al. (2008).
References:
1
Finlay et al. (2009);
2
Bourdôt et al. (2012);
3
Harris and Barker (2007);
4
Gallagher et al. (2012a);
5
Taylor, S. et al. (2012);
6
Sheppard (2012);
7
Sutherst et al.
(2000);
8
Laurance (2008);
9
Melloy et al. (2010);
10
Yeates and Newton (2009);
11
Watt et al. (2011b);
12
Thomson et al. (2010);
13
Gerard et al. (2012).
1398
Chapter 25 Australasia
25
Box 25-5 | Climate Change Vulnerability and Adaptation in Rural Areas
Rural communities in Australasia have higher proportions of older and unemployed people than urban populations (Mulet-Marquis
and Fairweather, 2008). Employment and economic prospects depend heavily on the physical environment and hence are highly
exposed to climate (averages, variability, and extremes) as well as changing commodity prices. These interact with other economic,
social, and environmental pressures, such as changing government policies (e.g., on drought, carbon pricing; Productivity Commission,
2009; Nelson et al., 2010) and access to water resources. The vulnerability of rural communities differs within and between countries,
reflecting differences in financial security, environmental awareness, policy and social support, strategic skills, and capacity for
diversification (Bi and Parton, 2008; Marshall, 2010; Nelson et al., 2010; Hogan et al., 2011b; Kenny, 2011).
Climate change will affect rural industries and communities through impacts on resource availability and distribution, particularly water.
Decreased availability and/or increased demand, or price, in response to climate change will increase tensions among agricultural,
mining, urban, and environmental water users (very high confidence), with implications for governance and participatory adaptation
processes to resolve conflicts (see Sections 25.4.2, 25.6.1, 25.7.2-3; Boxes 25-2, 25-10). Communities will also be affected through
direct impacts on primary production, extraction activities, critical infrastructure, population health, and recreational and culturally
significant sites (Kouvelis et al., 2010; Balston et al., 2012; see Sections 25.7-8).
Altered production and profitability risks and/or land use will translate into complex and interconnected effects on rural communities,
particularly income, employment, service provision, and reduced volunteerism (Stehlik et al., 2000; Bevin, 2007; Kerr and Zhang,
2009). The prolonged drought in Australia during the early 2000s, for example, had many interrelated negative social impacts in rural
communities, including farm closures, increased poverty, increased off-farm work, and, hence, involuntary separation of families,
increased social isolation, rising stress and associated health impacts, including suicide (especially of male farmers), accelerated rural
depopulation, and closure of key services (robust evidence, high agreement; Alston, 2007, 2010, 2012; Edwards and Gray, 2009;
Hanigan et al., 2012; see also Box CC-GC). Positive social change also occurred, however, including increased social capital through
interaction with community organizations (Edwards and Gray, 2009). While social and cultural changes have the potential to undermine
the adaptive capacity of communities (Smith, W. et al., 2011), robust ongoing engagement between farmers and the local community
can contribute to a strong sense of community and enhance potential for resilience (McManus et al., 2012; see also Section 25.4.3).
The economic impact of droughts on rural communities and the entire economy can be substantial. The most recent drought in
Australia (2006/7–2008/9), for example, is estimated to have reduced national GDP by about 0.75% (RBA, 2006) and regional GDP in
the southern Murray-Darling Basin was about 5.7% below forecast in 2007/08, along with the temporary loss of 6000 jobs (Wittwer
and Griffith, 2011). Widespread drought in New Zealand during 2007–2009 reduced direct and off-farm output by about NZ$3.6
billion (Butcher, 2009). The 2012–2013 drought in New Zealand is estimated to have reduced national GDP by 0.3 to 0.6% and
contributed to a significant rise in global dairy prices, which tempered even greater domestic economic losses (Kamber et al., 2013).
Drought frequency and severity are projected to increase in many parts of the region (Table 25-1).
The decisions of rural enterprise managers have significant consequences for and beyond rural communities (Pomeroy, 1996; Clark
and Tait, 2008). Many current responses are incremental, responding to existing climate variability (Kenny, 2011). Transformational
change has occurred where industries and individuals are relocating part of their operations in response to recent and/or expectations
of future climate or policy change (Kenny, 2011; see also Box 25-10), for example, rice (Gaydon et al., 2010), wine grapes (Park et al.,
2012), peanuts (Thorburn et al., 2012), or changing and diversifying land use in situ (e.g., the recent switch from grazing to cropping
in South Australia; Howden et al., 2010). Such transformational changes are expected to become more frequent and widespread with
a changing climate (high confidence; Section 25.7.2), with positive or negative implications for the wider communities in origin and
destination regions (Kiem and Austin, 2012).
Continued next page
1399
Australasia Chapter 25
25
Observed trends and modeling for wine grapes suggest that climate
change will lead to earlier budburst, ripening, and harvest for most
regions and scenarios (high confidence; Grace et al., 2009; Sadras and
Petrie, 2011; Webb, L.B. et al., 2012). Without adaptation, reduced quality
is expected in all Australian regions (high confidence; Webb et al., 2008).
Change in cultivar suitability in specific regions is expected (Clothier et
al., 2012), with potential for development of cooler or more elevated
sites within some regions (Tait, 2008; Hall and Jones, 2009) and/or
expansion to new regions, with some growers in Australia already
relocating (e.g., to Tasmania; Smart, 2010).
Climate change and elevated CO
2
impacts on weeds, pests, and diseases
are highly uncertain (see Box 25-4). Future performance of currently
effective plant resistance mechanisms under temperature and elevated
CO
2
is particularly important (Melloy et al., 2010; Chakraborty et al.,
2011), as is the future efficacy of widely used biocontrol—that is, the
introduction or stimulation of natural enemies to control pests (Gerard
et al., 2012). Australia is ranked second and New Zealand fourth in the
world in the number of biological control agent introductions (Cock et
al., 2010).
25.7.2.3. Integrated Adaptation Perspectives
Future water demand by the sector is critical for planning (Box 25-2).
Irrigated agriculture occupies less than 1% of agricultural land in
Australia but accounted for 28% of gross agricultural production value
in 2010–11; almost half of this was produced within the Murray-Darling
Basin, which used 68% of all irrigation water (ABS, 2012b; DAFF, 2012).
Reduced inflow under dry climate scenarios is predicted to reduce
substantially the value of agricultural production in the Basin (robust
evidence, high agreement; Garnaut, 2008; Quiggin et al., 2010; Qureshi
et al., 2013b)—for example, in one study by 12 to 44% to 2030 and 49
to 72% to 2050 (A1F1; Garnaut, 2008).
Water availability also constrains agricultural expansion: 17 Mha in
northern Australia could support cropping but only 1% has appropriate
water availability (Webster et al., 2009). In New Zealand, the irrigated
area has risen by 82% since 1999 to more than 1 Mha; 76% is on pasture
(Rajanayaka et al., 2010). The New Zealand dairy herd doubled between
1980–2009 expanding from high rainfall zones (>2000 mm annual) into
drier, irrigation-dependent areas (600 to 1000 mm annual); this
dependence will increase with further expansion (Robertson, 2010), which
is being supported by the Government’s Irrigation Acceleration Fund.
Many adaptation options—such as flexible water allocation, irrigation,
and seasonal forecasting—support managing risk in the current climate
(Howden et al., 2008; Botterill and Dovers, 2013) and adoption is often
high (Hogan et al., 2011a; Kenny, 2011).
However, incremental on-farm adaptation has limits (Park et al., 2012)
and may hinder transformational change such as diversification of
land use or relocation (see Box 25-5) if it encourages persistence where
climate change may take current systems beyond their response
capacity (Marshall, 2010; Park et al., 2012; Rickards and Howden, 2012).
In many cases, transformational change requires a greater level of
commitment, access to more resources, and greater integration across
all levels of decision making that encompass both on- and off-farm
knowledge, processes and values (Marshall, 2010; Rickards and Howden,
2012).
25.7.3. Mining
Australia is the world’s largest exporter of coking coal and iron ore and
has the world’s largest resources of brown coal, nickel, uranium, lead, and
zinc (ABS, 2012c). Recent events demonstrated significant vulnerability
to climate extremes: the 2011 floods reduced coal exports by 25 to 54
million tonnes and led to AU$5 to 9 billion revenue lost in that year
(ABARES, 2011; RBA, 2011), and tropical cyclones regularly disrupted
mining operations over the past decade (McBride, 2012; Sharma et al.,
2013). Flood impacts were exacerbated by regulatory constraints on mine
discharges, highlighting tensions among industry, social, and ecological
management objectives (QRC, 2011), and by flooding affecting road
and rail transport to major shipping ports (QRC, 2011; Sharma et al.,
2013).
Projected changes in climate extremes imply increasing sector
vulnerability without adaptation (high confidence; Hodgkinson et al.,
2010a,b). Stakeholders have conducted initial climate risk assessments
(Mills, 2009) and perceive the adaptive capacity of the industry to be
high (Hodgkinson et al., 2010a; Loechel et al., 2010; QRC, 2011), but
costs and broader benefits are yet to be explored along the value
chain and evaluated for community support. Ongoing challenges
include competition for energy and water, climate change skepticism,
dealing with contrasting extremes, avoiding maladaptation, and mining-
community relations regarding response options, acceptable mine
discharges, and post-mining rehabilitation (Loechel et al., 2013; Sharma
et al., 2013).
Box 25-5 (continued)
Although stakeholders within rural communities differ in their vulnerabilities and adaptive capacities, they are bound by similar
dependence on critical infrastructure and resources, economic conditions, government policy direction, and societal expectations
(Loechel et al., 2013). Consequently, adaptation to climate change will require an approach that devolves decision making to the
level where the knowledge for effective adaptations resides, using open communication, interaction, and joint planning (Nelson et
al., 2008; Kiem and Austin, 2013).
1400
Chapter 25 Australasia
25
25.7.4. Energy Supply, Demand, and Transmission
Energy demand is projected to grow by 0.5 to 1.3% per annum in
Australasia over the next few decades in the absence of major new
policies (MED, 2011; Syed, 2012). Australia’s predominantly thermal
power generation is vulnerable to drought-induced water restrictions,
which could require dry-cooling and increased water use efficiency
where rainfall declines (Graham et al., 2008; Smart and Aspinall, 2009).
Depending on carbon price and technology costs, renewable electricity
generation in Australia is projected to increase from 10% in 2010–11
to approximately 33 to 50% by 2030 (Hayward et al., 2011; Stark et al.,
2012; Syed, 2012), but few studies have explored the vulnerability
of these new energy sources to climate change (Bryan et al., 2010;
Crook et al., 2011; Odeh et al., 2011). New Zealand’s predominantly
hydroelectric power generation is vulnerable to precipitation variability.
Increasing winter precipitation and snow melt, and a shift from snowfall
to rainfall will reduce this vulnerability (medium confidence) as
winter/spring inflows to main hydro lakes are projected to increase by
5 to 10% over the next few decades (McKerchar and Mullan, 2004;
Poyck et al., 2011). Further reductions in seasonal snow and glacial melt
as glaciers diminish, however, would compromise this benefit (Chinn,
2001; Renwick et al., 2009; Srinivasan et al., 2011). Increasing wind
Box 25-6 | Climate Change and Fire
Fire during hot, dry, and windy summers in southern Australia can cause loss of life and substantial property damage (Cary et al.,
2003; Adams and Attiwill, 2011). The “Black Saturday” bushfires in Victoria in February 2009, for example, burned more than 3500
km
2
, caused 173 deaths, destroyed more than 2000 buildings, and caused damages of AU$4 billion (Cameron et al., 2009; VBRC,
2010). This fire occurred toward the end of a 13-year drought (CSIRO, 2010) and after an extended period of consecutive days over
30°C (Tolhurst, 2009).
Climate change is expected to increase the number of days with very high and extreme fire weather (Table 25-1), with greater
changes where fire is weather-constrained (most of southern Australia; many, in particular eastern and northern, parts of New
Zealand) than where it is constrained by fuel load and ignitions (tropical savannahs in Australia). Fire season length will be extended
in many already high-risk areas (high confidence) and so reduce opportunities for controlled burning (Lucas et al., 2007). Higher CO
2
may also enhance fuel loads by increasing vegetation productivity in some regions (Donohue et al., 2009; Williams et al., 2009;
Bradstock, 2010; Hovenden and Williams, 2010; King et al., 2011).
Climate change and fire will have complex impacts on vegetation communities and biodiversity (Williams et al., 2009). Greatest
impacts in Australia are expected in sclerophyll forests of the southeast and southwest (Williams et al., 2009). Most New Zealand
native ecosystems have limited exposure but also limited adaptations to fire (Ogden et al., 1998; McGlone and Walker, 2011). There is
high confidence that increased fire incidence will increase risk in southern Australia to people, property, and infrastructure such as
electricity transmission lines (Parsons Brinkerhoff, 2009; O'Neill and Handmer, 2012; Whittaker et al., 2013) and in parts of New
Zealand where urban margins expand into rural areas (Jakes et al., 2010; Jakes and Langer, 2012); exacerbate some respiratory
conditions such as asthma (Johnston et al., 2002; Beggs and Bennett, 2011); and increase economic risks to plantation forestry (Watt
et al., 2008; Pearce et al., 2011). Forest regeneration following wildfires also reduces water yields (Brown et al., 2005; MDBC, 2007),
while reduced vegetation cover increases erosion risk and material washoff to waterways with implications for water quality
(Shakesby et al., 2007; Wilkinson et al., 2009; Smith, H.G. et al., 2011).
In Australia, fire management will become increasingly challenging under climate change, potentially exacerbating conflicting
management objectives for biodiversity conservation versus protection of property (high confidence; O'Neill and Handmer, 2012;
Whittaker et al., 2013). Current initiatives center on planning and regulations, building design to reduce flammability, fuel management,
early warning systems, and fire detection and suppression (Handmer and Haynes, 2008; Preston et al., 2009; VBRC, 2010; O'Neill and
Handmer, 2012). Some Australian authorities are taking climate change into account when rethinking approaches to managing fire to
restore ecosystems while protecting human life and properties (Preston et al., 2009; Adams and Attiwill, 2011). Improved understanding
of climate drivers of fire risk is assisting fire management agencies, landowners, and communities in New Zealand (Pearce et al.,
2008, 2011), although changes in management to date show little evidence of being driven by climate change.
1401
Australasia Chapter 25
25
p
ower generation (MED, 2011) would benefit from projected increases
in mean westerly winds but face increased risk of damages and shutdown
during extreme winds (Renwick et al., 2009).
Climate warming would reduce annual average peak electricity demands
by 1 to 2% per degree Celsius across New Zealand and 2(±1)% in New
South Wales, but increase by 1.1(±1.4)% and 4.6(±2.7)% in Queensland
and South Australia due to air conditioning demand (Stroombergen et
al., 2006; Jollands et al., 2007; Thatcher, 2007; Nguyen et al., 2010).
Increased summer peak demand, particularly in Australia (see also
Figure 25-5), will place additional stress on networks and can result in
blackouts (very high confidence; Jollands et al., 2007; Thatcher, 2007;
Howden and Crimp, 2008; Wang et al., 2010a). During the 2009 Victorian
heat wave, demand rose by 24% but electrical losses from transmission
lines increased by 53% due to higher peak currents (Nguyen et al.,
2010), and successive failures of the overloaded network temporarily
left more than 500,000 people without power (QUT, 2010). Various
adaptation options to limit increasing urban energy demand exist and
some are being implemented (see Box 25-9).
There is limited evidence but high agreement that without additional
adaptation, distribution networks in most Australian states will be at
high risk of failure by 2031–2070 under non-mitigation scenarios due
to increased bushfire risk and potential strengthening and southward
shift of severe cyclones in tropical regions (Maunsell and CSIRO, 2008;
Parsons Brinkerhoff, 2009). Adaptation costs have been estimated at
AU$2.5 billion to 2015, with more than half to meet increasing demand
for air conditioning and the remainder to increase resilience to climate-
related hazards; underground cabling would reduce bushfire risk but
has large investment costs that are not included (Parsons Brinkerhoff,
2009). Decentralized ownership of assets constitutes a significant
adaptation constraint (ATSE, 2008; Parsons Brinkerhoff, 2009). In New
Zealand, increasing high winds and temperatures have been identified
qualitatively as the most relevant risks to transmission (Jollands et al.,
2007; Renwick et al., 2009).
25.7.5. Tourism
Tourism contributes 2.6 to 4% of GDP to the economies of Australia
and New Zealand (ABS, 2010a; SNZ, 2011). The net present value of the
Great Barrier Reef alone over the next 100 years has been estimated at
AU$51.4 billion (Oxford Economics, 2009). Most Australasian tourism is
exposed to climate variability and change (see Section 25.2 for projected
trends), and some destinations are highly sensitive to extreme events
(Hopkins et al., 2012). The 2011 floods and Tropical Cyclone Yasi, for
example, cost the Queensland tourism industry about AU$590 million,
mainly due to cancellations and damage to the Great Barrier Reef (PwC,
2011); and drought in the Murray-Darling Basin caused an estimated
AU$70 million loss in 2008 due to reduced visitor days (TRA, 2010).
25.7.5.1. Projected Impacts
Future impacts on tourism have been modeled for several Australian
destinations. The Great Barrier Reef is expected to degrade under all
climate change scenarios (Sections 25.6.2, 30.5; Box CC-CR), reducing
i
ts attractiveness (Marshall and Johnson, 2007; Bohensky et al., 2011;
Wilson and Turton, 2011b). Ski tourism is expected to decline in the
Australian Alps due to snow cover reducing more rapidly than in New
Zealand (Pickering et al., 2010; Hendrikx et al., 2013) and greater
perceived attractiveness of New Zealand (Hopkins et al., 2012). Higher
temperature extremes in the Northern Territory are projected with
high confidence to increase heat stress and incur higher costs for air
conditioning (Turton et al., 2009). Sea level rise places pressures on
shorelines and long-lived infrastructure but implications for tourist
resorts have not been quantified (Buckley, 2008).
Economic modeling suggests that the Australian alpine region would
be most negatively affected in relative terms due to limited alternative
activities (Pham et al., 2010), whereas the competitiveness of some
destinations (e.g., Margaret River in Western Australia) could be enhanced
by higher temperatures and lower rainfall (Jones et al., 2010; Pham et
al., 2010). An analog-based study suggests that, in New Zealand,
warmer and drier conditions mostly benefit but wetter conditions and
extreme climate events undermine tourism (Wilson and Becken, 2011).
Confidence in outcomes is low, however, owing to uncertain future
tourist behaviour (Scott et al., 2012; see also Section 25.9.2).
25.7.5.2. Adaptation
Both New Zealand and Australia have formalized adaptation strategies
for tourism (Becken and Clapcott, 2011; Zeppel and Beaumont, 2011).
In Australia, institutions at various levels also promote preparation for
extreme events (Tourism Queensland, 2007, 2010; Tourism Victoria,
2010) and strengthening ecosystem resilience to maintain destination
attractiveness (GBRMPA, 2009b). Snow-making is already broadly
adopted to increase reliability of skiing (Bicknell and McManus, 2006;
Hennessy et al., 2008b), but its future effectiveness depends on location.
In New Zealand, even though warming will significantly reduce the
number of days suitable for snow-making (Hendrikx and Hreinsson, 2012),
sufficient snow could be made in all years until the end of the 21st century
to maintain current minimum operational skiing conditions. Options for
resorts in Australia’s Snowy Mountains are far more limited (Hendrikx et
al., 2013), where maintaining skiing conditions until at least 2020 would
require AU$100 million in capital investment into 700 snow guns and 2.5
to 3.3 GL of water per month (Pickering and Buckley, 2010).
Short investment horizons, high substitutability, and a high proportion
of human capital compared with built assets give high confidence that
the adaptive capacity of the tourism industry is high overall, except for
destinations where climate change is projected to degrade core natural
assets and diversification opportunities are limited (Evans et al., 2011;
Morrison and Pickering, 2011). Strategic adaptation decisions are
constrained by uncertainties in regional climatic changes (Turton et al.,
2010), limited concern (Bicknell and McManus, 2006), lack of leadership,
and limited coordinated forward planning (Sanders et al., 2008; Turton
et al., 2009; Roman et al., 2010; White and Buultjens, 2012). An
integrated assessment of tourism vulnerability in Australasia is not yet
possible owing to limited understanding of future changes in tourism
and community preferences (Scott et al., 2012), including the flow-on
effects of changing travel behavior and tourism preferences in other
world regions (see Section 25.9.2).
1402
Chapter 25 Australasia
25
25.8. Human Society
2
5.8.1. Human Health
25.8.1.1. Observed Impacts
Life expectancy in Australasia is high, but shows substantial ethnic and
socioeconomic inequalities (Anderson et al., 2006). Mortality increases
in hot weather in Australia (robust evidence, high agreement; Bi and
Parton, 2008; Vaneckova et al., 2008) with air pollution exacerbating
this association. The last 4 decades have seen a steady increase in the
ratio of summer to winter mortality in Australia, indicating a health
effect from climatic warming (Bennett et al., 2013). Exceptional heat
wave conditions in Australia have been associated with substantial
increases in mortality and hospital admissions in several regional towns
and capital cities (high confidence; Khalaj et al., 2010; Loughnan et al.,
2010; Tong et al., 2010a,b). For example, during the heat wave in
January and February 2009 in southeastern Australia (BoM, 2009), total
e
mergency cases increased by 46% over the three hottest days: direct
heat-related health problems increased 34-fold, 61% of these being in
people aged 75 years or older, and there were an estimated 374 excess
deaths, a 62% increase in all-cause mortality (Victorian Government,
2009a). Mental health admissions increased across all age groups by
7.3% in metropolitan South Australia during heat waves (1993–2006;
Hansen et al., 2008). Mortality attributed to mental and behavioral
disorders increased in the 65- to 74-year age group and in persons with
existing mental health problems (Hansen et al., 2008). Experience of
extreme events also strongly affects psychological well-being (see
Section 25.4.3).
25.8.1.2. Projected Impacts
Projected increases in heat waves (Figure 25-5) will increase heat-related
deaths and hospitalizations, especially among the elderly, compounded
by population growth and aging (high confidence; Bambrick et al., 2008;
20
40
60
80
100
140
120
New South
Wales
Victoria Queensland South
Australia
Western
Australia
Tasmania Northern
Territories
Australian
Capital Territory
~1990
~2050
~2100
Person-days above 40°C (millions)
Days per year > 40°C
0–2
2–5
5–10
10–20
20–40
40–60
60–90
90–130
130–192
140
Days per year > 40
°
C
Figure 25-5 | Projected changes in exposure to heat under a high emissions scenario (A1FI). Maps show the average number of days with peak temperatures >40°C, for
approximately 1990 (based on available meteorological station data for the period 1975–2004), approximately 2050, and approximately 2100. Bar charts show the change in
population heat exposure, expressed as person-days exposed to peak temperatures >40°C, aggregated by State/Territory and including projected population growth for a default
scenario. Future temperatures are based on simulations by the Geophysical Fluid Dynamics Laboratory Coupled Model version 2 (GFDL-CM2) global climate model (Meehl et al.,
2007), re-scaled to the A1FI scenario; simulations based on other climate models could give higher or lower results. Data from Baynes et al. (2012).
~1990 ~2050 ~2100
1403
Australasia Chapter 25
25
Gosling et al., 2009; Huang et al., 2012). In the southern states of
Australia and parts of New Zealand, this may be partly offset by reduced
deaths from cold at least for modest rises in temperature (low confidence;
Bambrick et al., 2008; Kinney, 2012). With strong mitigation, climate
change is projected to result in 11% fewer temperature-related deaths
in both 2050 and 2100 in Australia, but 14% and 100% more deaths in
2050 and 2100, respectively, without mitigation under a hot, dry A1FI
scenario (Bambrick et al., 2008; see Chapter 11 for detail on temperature-
related health trade-offs). Net results were driven almost entirely by
increased mortality in the north, especially Queensland, consistent with
Huang et al. (2012). In a separate study that accounted for increased daily
temperature variability, a threefold increase in heat-related deaths is
projected for Sydney by 2100 for the A2 scenario, assuming no adaptation
(Gosling et al., 2009). The number of hot days when physical labor in
the sun becomes dangerous is also projected to increase substantially
in Australia by 2070, leading to economic costs from lost productivity,
increased hospitalizations, and occasional deaths (medium confidence;
Hanna et al., 2011; Maloney and Forbes, 2011).
Water- and food-borne diseases are projected to increase, but the
complexity of their relationship to climate and non-climate drivers
means there is low confidence in specific projections. For Australia,
205,000 to 335,000 new cases of bacterial gastroenteritis by 2050, and
239,000 to 870,000 cases by 2100, are projected under a range of
emission scenarios (Bambrick et al., 2008; Harley et al., 2011). Based
on their observed positive relationship with temperature, notifications
of salmonellosis notifications are projected to increase 15% for every
1°C increase in average monthly temperatures (Britton et al., 2010a).
Water-borne zoonotic diseases such as cryptosporidiosis and giardiasis
have more complex relationships with climate and are amenable to
various adaptations, making future projections more difficult (Britton et
al., 2010b; Lal et al., 2012).
Understanding the combined effects of climate change and socioeconomic
development on the distribution of vector-borne diseases has improved
since the AR4. Australasia is projected to remain malaria free under the
A1B emission scenario until at least 2050 (Béguin et al., 2011) and
Box 25-7 | Insurance as a Climate Risk Management Tool
Insurance helps spread the risk from extreme events across communities and over time and therefore enhances the resilience of
society to disasters (see Section 10.7). In Australia, insured losses are dominated by meteorological hazards, including the 2011
Queensland floods and the 1999 Sydney hailstorm (ICA, 2012) with estimated claims of AU$3 billion per annum (IAA, 2011b). In New
Zealand, floods and storms are the second most costly natural hazards after earthquakes (ICNZ, 2013). The number of damaging
insured events (up to a certain loss value) has increased significantly in the Oceania region since 1980 (Schuster, 2013). Normalized
losses in Australia show no significant trend from at least 1967 to 2006 (Crompton and McAneney, 2008; Crompton et al., 2010;
Table 10-4), consistent with the global conclusion (IPCC, 2012) that increasing exposure of people and economic assets has been the
major cause of long-term increases in economic losses from weather- and climate-related disasters. Issues relating to data quality
and methodological choices prevent definitive conclusions regarding the role of climate change in loss trends (Crompton et al., 2011;
Nicholls, 2011; IPCC, 2012).
There is high confidence that, without adaptive measures, projected increases in extremes (Table 25-1) and uncertainties in these
projections will lead to increased insurance premiums, exclusions, and non-coverage in some locations (IAG, 2011), which will
reshape the distribution of vulnerability, for example, through unaffordability or unavailability of cover in areas at highest risk (IAA,
2011a,b; NDIR, 2011; Booth and Williams, 2012). Restriction of cover occurred in some locations following the 2011 flood events in
Queensland (Suncorp, 2013).
Insurance can contribute positively to risk reduction by providing incentives to policy holders to reduce their risk profile (O'Neill and
Handmer, 2012), for example, through resilience ratings given to buildings (TGA, 2009; Edge Environment, 2011; IAG, 2011). Apart
from constituting an autonomous private sector response to extreme events, insurance can also be framed as a form of social policy
to manage climate risks, similar to New Zealand’s government insurance scheme (Glavovic et al., 2010); government measures to
reduce or avoid risks also interact with insurance companies’ willingness to provide cover (Booth and Williams, 2012). Yet insurance
can also act as a constraint on adaptation, if those living in climate-risk prone localities pay discounted or cross-subsidized premiums
or policies fail to encourage betterment after damaging events by requiring replacement of “like for like,” constituting a missed
opportunity for risk reduction (NDIR, 2011; QFCI, 2012; Reisinger et al., 2013; see also Section 10.7). The effectiveness of insurance
thus depends on the extent to which it is linked to a broader national resilience approach to disaster mitigation and response
(Mortimer et al., 2011).
1404
Chapter 25 Australasia
25
sporadic cases could be treated effectively. The area climatically suitable
for transmission of dengue will expand in Australasia (high confidence;
Bambrick et al., 2008; Åström et al., 2012), but changes in socioeconomic
factors, especially domestic water storage, may have a more important
influence on disease incidence than climate (Beebe et al., 2009; Kearney
et al., 2009). Impacts of climate change on Barmah Forest Virus in
Queensland depend on complex interactions between rainfall and
temperature changes, together with tidal and socioeconomic factors,
and thus will vary substantially among different coastal regions (Naish
et al., 2013). The effects of climate change combined with frequent
Box 25-8 | Changes in Flood Risk and Management Responses
Flood damages across eastern Australia and both main islands of New Zealand in 2010 and 2011 revealed a significant adaptation
deficit (ICA, 2012; ICNZ, 2013). For example, the Queensland floods in January 2011 resulted in 35 deaths, three-quarters of the State
including Brisbane declared a disaster zone, and damages to public infrastructure of AU$5 to 6 billion (Queensland Government,
2011). These floods were associated with a strong monsoon and the strongest La Niña on record (Cai et al., 2012; CSIRO and BoM,
2012; Evans and Boyer-Souchet, 2012). Flood frequency and severity exhibit strong decadal variability with no significant long-term
trend in Australasia to date (Kiem et al., 2003; Smart and McKerchar, 2010; Ishak et al., 2013).
Flood risk is projected to increase in many regions due to more intense extreme rainfall events driven by a warmer and wetter
atmosphere (medium confidence; Table 25-1). High-resolution downscaling (Carey-Smith et al., 2010), and dynamic catchment
hydrological and river hydraulic modeling in New Zealand (Gray, W. et al., 2005; McMillan et al., 2010; MfE, 2010b; Ballinger et al.,
2011; Duncan and Smart, 2011; McMillan et al., 2012) indicate that the 50-year and 100-year flood peaks for rivers in many parts of
the country will increase by 5 to 10% by 2050 and more by 2100 (with large variation between models and emissions scenarios),
with a corresponding decrease in return periods for specific flood levels. Studies for Queensland show similar results (DERM et al.,
2010). In Australia, flood risk is expected to increase more in the north (driven by convective rainfall systems) than in the south
(where more intense extreme rainfall may be compensated by drier antecedent moisture conditions), consistent with confidence in
heavy rainfall projections (Table 25-1; Alexander and Arblaster, 2009; Rafter and Abbs, 2009).
Flood risk near river mouths will be exacerbated by storm surge associated with higher sea level and potential change in wind
speeds (McInnes et al., 2005; MfE, 2010b; Wang et al., 2010b). Higher rainfall intensity and peak flow will also increase erosion and
sediment loads in waterways (Prosser et al., 2001; Nearing et al., 2004) and exacerbate problems from aging stormwater and
wastewater infrastructure in cities (Howe et al., 2005; Jollands et al., 2007; CCC, 2010; WCC, 2010; see also Box 25-9). However,
moderate flooding also has benefits through filling reservoirs, recharging groundwater, and replenishing natural environments
(Hughes, 2003; Chiew and Prosser, 2011; Oliver and Webster, 2011).
Adaptation to increased flood risk from climate change is starting to happen (Wilby and Keenan, 2012) through updating guidelines
for design flood estimation (MfE, 2010b; Westra, 2012), improving flood risk management (O'Connell and Hargreaves, 2004; NFRAG,
2008; Queensland Government, 2011), accommodating risk in flood prone areas (options include raising floor levels, using strong
piled foundations, using water-resistant insulation materials, and ensuring weather tightness), and risk reduction and avoidance
through spatial planning and managed relocation (Trotman, 2008; Glavovic et al., 2010; LVRC, 2012; QFCI, 2012). Adaptation options
in urban areas also include ecosystem-based approaches such as retaining floodplains and floodways, restoring wetlands, and
retrofitting existing systems to attenuate flows (Howe et al., 2005; Skinner, 2010; WCC, 2010; see also Box 25-9).
The recent flooding in eastern Australia and the projected increase in future flood risk have resulted in changes to reservoir operations
to mitigate floods (van den Honert and McAneney, 2011; QFCI, 2012) and insurance practice to cover flood damages (NDIR, 2011;
Phelan, 2011; see also Box 25-7). However, the magnitude of potential future changes in flood risk and limits to incremental
adaptation responses in urban areas suggest that more transformative approaches based on altering land use and avoidance of
exposure to future flooding may be needed in some locations, especially if changes in the upper range of projections are realized (high
confidence; Lawrence and Allan, 2009; DERM et al., 2010; Glavovic et al., 2010; Wilby and Keenan, 2012; Lawrence et al., 2013a).
1405
Australasia Chapter 25
25
t
ravel within and outside the region, and recent incursions of exotic
mosquito species, could expand the geographic range of other important
arboviruses such as Ross River Virus (medium confidence; Derraik and
Slaney, 2007; Derraik et al., 2010).
A growing literature since the AR4 has focused on the psychological
impacts of climate change, based on impacts of recent climate variability
and extremes (Doherty and Clayton, 2011; see also Section 25.4.3).
These studies indicate significant mental health risks associated with
climate-related disasters, in particular persistent and severe drought,
floods, and storms; climate impacts may be especially acute in
rural communities where climate change places additional stresses on
livelihoods (high confidence; Edwards et al., 2011; see also Box 25-5).
Projected population growth and urbanization could further increase
health risks indirectly via climate-related stress on housing, transport
and energy infrastructure, and water supplies (low confidence; Howden-
Chapman, 2010; see also Box 25-9).
25.8.1.3. Adaptation
Research since the AR4 has mainly focused on climate change impacts,
although some adaptation strategies have received attention in Australia.
These include improving health care services, social support for those most
at risk, improving community awareness to reduce adverse exposures,
developing early warning and emergency response plans (Wang and
McAllister, 2011), and understanding perceptions of climatic risks to
health as they affect adaptive behaviors (Akompad et al., 2013). In New
Zealand, central Government health policies do not identify specific
measures to adapt to climate change (Wilson, 2011). In both countries,
policies to reduce risks from extreme events such as floods and fires
will have co-benefits for health (see Boxes 25-6, 25-8).
A review of the southern Australian heat wave of 2009 identified a
range of issues including communication failures with no clear public
information or warning strategy, and no clear thresholds for initiating
public information campaigns (Kiem et al., 2010). Emergency services
were underprepared and relied on reactive solutions (QUT, 2010).
The Victorian government has since developed a heat wave plan to
coordinate a state-wide response, maintain consistent community-wide
understanding through a Heat Health alert system, build capacity of
councils to support communities most at risk, support a Heat Health
Intelligence surveillance system, and distribute public health information
(Victorian Government, 2009b).
25.8.2. Indigenous Peoples
25.8.2.1. Aboriginal and Torres Strait Islanders
Work since the AR4 includes a national Indigenous adaptation research
action plan (Langton et al., 2012), regional risk studies (Green et al.,
2009; DNP, 2010; TSRA, 2010; Nursey-Bray et al., 2013) and scrutiny
from an Indigenous rights perspective (ATSISJC, 2009). Socioeconomic
disadvantage and poor health (SCRGSP, 2011) indicate a disproportionate
climate change vulnerability of Indigenous Australians (McMichael et al.,
2009) although there are no detailed assessments. In urban and regional
areas, where 75% of the Indigenous population lives (ABS, 2010b),
assessments have not specifically addressed risks to Indigenous people
(e.g., Guillaume et al., 2010). In other regions, all remote, there is limited
empirical evidence of vulnerability (Maru et al., 2012). However, there
is medium evidence and high agreement for significant future impacts
from increasing heat stress, extreme events, and increased disease
(Campbell et al., 2008; Spickett et al., 2008; Green et al., 2009).
The Indigenous estate comprises more than 25% of the Australian land
area (Altman et al., 2007; NNTR, 2013). There is high agreement but
limited evidence that natural resource dependence (e.g., Bird et al.,
2005; Gray, M.C. et al., 2005; Kwan et al., 2006; Buultjens et al., 2010)
increases Indigenous exposure and sensitivity to climate change (Green
et al., 2009); climate change-induced dislocation, attenuation of cultural
attachment to place, and loss of agency will disadvantage Indigenous
mental health and community identity (Fritze et al., 2008; Hunter, 2009;
McIntyre-Tamwoy and Buhrich, 2011); and, housing, infrastructure,
services, and transport, often already inadequate for Indigenous needs
especially in remote Australia (ABS, 2010c), will be further stressed
(Taylor and Philp, 2010). Torres Strait island communities and livelihoods
are vulnerable to major impacts from even small sea level rises (high
confidence; DCC, 2009; Green, D. et al., 2010a; TSRA 2010).
Little adaptation of Indigenous communities to climate change is
apparent to date (cf. Burroughs, 2010; GETF 2011; Nursey-Bray et al.,
2013; Zander et al., 2013). Plans and policies that are imposed on
Indigenous communities can constrain their adaptive capacity (Ellemor,
2005; Petheram et al., 2010; Veland et al., 2010; Langton et al., 2012)
but participatory development of adaptation strategies is challenged
by multiple stressors and uncertainty about causes of observed changes
(Leonard, S. et al., 2010; Nursey-Bray et al., 2013). Adaptation planning
would benefit from a robust typology (Maru et al., 2011) across the
diversity of Indigenous life experience (McMichael et al., 2009). Indigenous
re-engagement with environmental management (e.g., Hunt et al.,
2009; Ross et al., 2009) can promote health (Burgess et al., 2009) and
may increase adaptive capacity (Berry et al., 2010; Davies et al., 2011).
There is emerging interest in integrating Indigenous observations of
climate change (Green, D. et al., 2010b; Petheram et al., 2010) and
developing inter-cultural communication tools (Leonard, S. et al., 2010;
Woodward et al., 2012). Extensive land ownership in northern and
inland Australia and land management traditions mean that Indigenous
people are well situated to provide greenhouse gas abatement and
carbon sequestration services that may also support their livelihood
aspirations (Whitehead et al., 2009; Heckbert et al., 2012).
25.8.2.2. New Zealand Māori
The projected impacts of climate change on Māori society are expected
to be highly differentiated, reflecting complex economic, social, cultural,
environmental, and political factors (high confidence). Since the AR4,
studies have been either sector-specific (e.g., Insley, 2007; Insley and
Meade, 2008; Harmsworth et al., 2010; King et al., 2012) or more general,
inferring risk and vulnerability based on exploratory engagements with
varied stakeholders and existing social, economic, political, and ecological
conditions (e.g., MfE, 2007b; Te Aho, 2007; King et al., 2010).
1406
Chapter 25 Australasia
25
T
he Māori economy depends on climate-sensitive primary industries
with vulnerabilities to climate conditions (high confidence; Packman et
al., 2001; NZIER, 2003; Cottrell et al., 2004; TPK, 2007; Tait et al., 2008b;
Harmsworth et al., 2010; King et al., 2010; Nana et al., 2011a). Much of
Māori-owned land is steep (>60%) and susceptible to damage from
high intensity rainstorms, while many lowland areas are vulnerable to
flooding and sedimentation (Harmsworth and Raynor, 2005; King et al.,
2010). Land in the east and north is also drought prone, and this increases
uncertainties for future agricultural performance, product quality, and
investment (medium confidence; Cottrell et al., 2004; Harmsworth et
al., 2010; King et al., 2010). The fisheries and aquaculture sector faces
substantial risks (and uncertainties) from changes in ocean temperature
and chemistry, potential changes in species composition, condition, and
productivity levels (medium confidence; King et al., 2010; see also
Section 25.6.2). At the community and individual level, Māori regularly
utilize the natural environment for hunting and fishing, recreation, the
maintenance of traditional skills and identity, and collection of cultural
resources (King and Penny, 2006; King et al., 2012). Many of these
activities are already compromised due to resource competition,
degradation, and modification (Woodward et al., 2001; King et al.,
2012). Climate change driven shifts in natural ecosystems will further
challenge the capacities of some Māori to cope and adapt (medium
confidence; King et al., 2012).
Māori organizations have sophisticated business structures, governance
(e.g., trusts, incorporations), and networks (e.g., Iwi leadership groups)
across the state and private sectors (Harmsworth et al., 2010; Insley, 2010;
Nana et al., 2011b), critical for managing and adapting to climate change
risks (Harmsworth et al., 2010; King et al., 2012). Future opportunities
will depend on partnerships in business, science, research, and government
(high confidence; Harmsworth et al., 2010; King et al., 2010) as well as
innovative technologies and new land management practices to better
suit future climates and use opportunities from climate policy, especially
in forestry (Carswell et al., 2002; Harmsworth, 2003; Funk and Kerr,
2007; Insley and Meade, 2008; Tait et al., 2008b; Penny and King,
2010).
M
āori knowledge of environmental processes and hazards (King et al.,
2005, 2007) as well as strong social-cultural networks are vital for
adaptation and ongoing risk management (King et al., 2008); however,
choices and actions continue to be constrained by insufficient resourcing,
shortages in social capital, and competing values (King et al., 2012).
Combining traditional ways and knowledge with new and untried policies
and strategies will be key to the long-term sustainability of climate-
sensitive Māori communities, groups, and activities (high confidence;
Harmsworth et al., 2010; King et al., 2012).
25.9. Interactions among Impacts,
Adaptation, and Mitigation Responses
The AR4 found that individual adaptation responses can entail synergies
or trade-offs with other adaptation responses and with mitigation, but
that integrated assessment tools were lacking in Australasia (Hennessy
et al., 2007). Subsequent studies provide detail on such interactions and
can inform a balanced portfolio of climate change responses, but
evaluation tools remain limited, especially for local decision making
(Park et al., 2011). A review of 25 specific climate change-associated
land use plans from Australia, for example, found that 14 exhibited
potential for conflict between mitigation and adaptation (Hamin and
Gurran, 2009).
25.9.1. Interactions among Local-Level Impacts,
Adaptation, and Mitigation Responses
Table 25-6 shows examples of adaptation responses that are either
synergistic or entail trade-offs with other impacts and/or adaptation
responses and goals. Adapting proactively to projected climate changes,
particularly extremes such as floods or drought, can increase near-
term resilience to climate variability and be a motivation for adopting
adaptation measures (Productivity Commission, 2012). However,
exclusive reliance on near-term benefits can increase trade-offs and
Box 25-9 | Opportunities, Constraints, and Challenges to Adaptation in Urban Areas
Considerable opportunities exist for Australasian cities and towns to reduce climate change impacts and, in some regions, benefit
from projected changes such as warmer winters and more secure water supply (Fitzharris, 2010; Australian Government, 2012). Many
tools and practices developed for sustainable resource management or disaster risk reduction in urban areas are co-beneficial for
climate change adaptation, and vice versa, and can be integrated with mitigation objectives (Hamin and Gurran, 2009). Despite the
abundance of potential adaptation options, however, social, cultural, institutional, and economic factors frequently constrain their
implementation (high confidence; see also Section 25.4.2). The form and longevity of cities and towns, with their concentration of
hard and critical infrastructure such as housing, transport, energy, stormwater and wastewater systems, telecommunications, and
public facilities provide additional challenges (see also Chapters 8, 10; Sections 25.7.4, 25.8.1; Boxes 25-1, 25-2, 25-8). Transport
infrastructure is vulnerable to extreme heat and flooding (QUT, 2010; Taylor and Philp, 2010) but quantification of future risks
remains limited (Gardiner et al., 2009; Balston et al., 2012; Baynes et al., 2012). Table 25-5 summarizes some adaptation options,
co-benefits, and constraints on their adoption in Australasia.
Continued next page
1407
Australasia Chapter 25
25
result in long-term maladaptation (high confidence). For example,
enhancing protection measures after major flood events, combined with
rapid re-building, accumulates fixed assets that can become increasingly
costly to protect as climate change continues, with attendant loss of
amenity and environmental values (Glavovic et al., 2010; Gorddard et
al., 2012; McDonald, 2013). Similarly, deferring adoption of increased
design wind speeds in cyclone-prone areas delays near-term investment
costs but also reduces the long-term benefit/cost ratio of the strategy
(Stewart and Wang, 2011).
Mitigation actions can contribute to but also counteract local adaptation
goals. Energy-efficient buildings, for example, reduce network and health
risks during heat waves, but urban densification to reduce transport
energy demand intensifies urban heat islands and, hence, heat-related
health risks (Sections 25.7.4, 25.8.1). Specific adaptations can also make
achievement of mitigation targets harder or easier. Increased use of
air conditioning, for example, increases energy demand, but energy
efficiency and building design can reduce heat exposure as well as
energy demand (Section 25.7.4, Box 25-9). Table 25-7 gives further
Box 25-9 (continued)
Overall, the implementation of climate change adaptation policy for urban settlements in Australia and New Zealand has been mixed.
The Australian National Urban Policy encourages adaptation, and many urban plans include significant adaptation policies (e.g., City
of Melbourne, 2009; City of Port Phillip, 2010; ACT Government, 2012; City of Adelaide, 2012). New Zealand also promotes urban
adaptation through strategies, plans, and guidance documents (MfE, 2008b; CCC, 2010; WCC, 2010; Auckland Council, 2012; NIWA et
al., 2012). Many examples of incremental urban adaptation exist (Box 25-2; Table 25-5), particularly where these include co-benefits
and respond to other stressors, like prolonged drought in southern Australia and recurrent floods. Experience is much scarcer with
more flexible land uses, managed relocation, and ecosystem-based adaptation that could transform existing settlement patterns and
development trends, and where maintaining flexibility to address long-term climate risks can run against near-term development
pressures (see Boxes 25-1, 25-2, 25-8, CC-EA). Decision-making models that support such adaptive and transformative changes
(Section 25.4.2; Box 25-1) have not yet been implemented widely in urban contexts; increased coordination among different levels of
government may be required to spread costs and balance public and private, near- and long-term, and local and regional benefits
(Norman, 2009, 2010; Britton, 2010;Abel et al., 2011; Lawrence et al., 2013a; McDonald, 2013; Palutikof et al., 2013; Reisinger et al., 2013).
Climate impact Adaptation options Co-benefi ts Barriers to adoption
Hot days and heat
w
aves
1 8
Greening cities / roofs; more green spaces;
w
ell-designed energy effi cient buildings;
occupant behavioral change; standards for new
a
nd retrofi tting of existing infrastructure and
assets; new methods and material for transport
infrastructure to withstand higher extreme
temperature
Energy effi ciency; reduced risk of blackouts;
f
ewer health impacts; resilient infrastructure and
assets; resilient community
Lack of standards; high installation costs; limited
u
nderstanding of benefi ts; high individual
discount rate; split of private costs and public
b
enefi ts
Decreased water supply
and drought (see Box
2
5-2 for more)
Supply augmentation (water recycling, rainwater
harvesting, increased storage, desalinization);
d
emand management; infrastructure upgrades;
integrated water-sensitive urban design
Water self-suffi ciency for current and future
demand / population; less pipe /storage leakage;
r
educed environmental impacts from abstraction
Potential health impacts of recycled water;
lower than expected uptake of demand options
a
nd relaxation after crises; trade-offs between
supply and demand management; cost and
e
nvironmental impacts of some augmentation
options
River and local fl ooding,
coastal erosion and
i
nundation (see Boxes
25-1 and 25-8 for more)
New standards and improvements to
building, water infrastructure (e.g., drainage
a
nd sewerage) and transport infrastructure;
upgrades of protection systems; retaining
oodplains / oodways; restoring wetlands;
buffers from hazard-prone areas; raising
minimum fl oor levels; rezoning / relocation
Reduced damages to homes and infrastructure
and loss of life; decreased insurance premiums;
h
abitat protection
High implementation cost especially
if retrospective on existing stock;
r
ezoning / relocation can affect property prices
and are highly contested.
Severe storms and
t
ropical cyclones
9 12
New building design to withstand higher wind
p
ressures; rezoning / relocation
Reduced damages to homes and infrastructure
a
nd loss of life; decreased insurance premiums
High implementation cost; rezoning / relocation
c
an affect property prices and are highly
contested.
Corrosion from
increased atmospheric
CO
2
levels
13,14
Improved standards for construction using
concrete; application of coatings for existing
building stock
Reduced rates of carbonation-induced corrosion
of concrete
Effectiveness of coatings varies with age and
condition of concrete.
T
able 25-5 |
E
xamples of co-benefi cial climate change adaptation options for urban areas and barriers to their adoption. Options in italics are already widely
implemented in Australia and New Zealand urban areas.
References:
1
BRANZ (2007);
2
Coutts et al. (2010);
3
Moon and Han (2011);
4
Stephenson et al. (2010);
5
Williams et al. (2010);
6
CSIRO et al. (2007);
7
Taylor and Philp
(2010);
8
QUT (2010);
9
Mason and Haynes (2010);
1
0
Wang et al. (2010b);
1
1
Stewart and Wang (2011);
1
2
Mason et al. (2013);
1
3
Stewart et al. (2012);
1
4
Wang et al. (2012).
1408
Chapter 25 Australasia
25
examples, and Box 25-10 explores the multiple and complex benefits
and trade-offs in changing land use to simultaneously adapt to and
mitigate climate change.
25.9.2. Intra- and Inter-regional Flow-on Effects
among Impacts, Adaptation, and Mitigation
Recent studies strengthen conclusions from the AR4 (Hennessy et al.,
2007) that flow-on effects from climate change impacts occurring in
other world regions can exacerbate or counteract projected impacts
in Australasia. Modeling suggests Australia’s terms of trade would
deteriorate by about 0.23% in 2050 and 2.95% in 2100 as climate
change impacts without mitigation reduce economic activity and
demand for coal, minerals, and agricultural products in other world
regions (A1FI scenario; Harman et al., 2008). As a result, Australian
Gross National Product (GNP) is expected to decline more strongly than
GDP because of climate change, especially toward the end of the 21st
century (Gunasekera et al., 2008). These conclusions, however, merit
only medium confidence, because they rely on simplified assumptions
Primary goal Sector(s) affected Examples of interactions between impacts and adaptation responses
Reduction of bushfi re risk in
n
atural landscapes
Biodiversity, tourism Potential for greater confl ict between conservation managers and other park users in Kosciuszko National Park if increasing fi re
i
ncidence causes park closures, either to reduce risk, or to rehabilitate vegetation after fi res (Wyborn, 2009), e.g., objectives of the
W
ildfi re Management Overlay (WMO) in Victoria confl ict with vegetation conservation (Hughes and Mercer, 2009).
Reduction of risk to energy
t
ransmission from bushfi res
Biodiversity, energy Underground cabling would reduce both the susceptibility of transmission networks to fi re and ignition sources for wild fi res, thus
r
educing risks to ecosystems and settlements; constraints include signifi cant investment cost, diverse ownership of assets, and
lack of an overarching national strategy (ATSE, 2008; Parsons Brinkerhoff, 2009; Linnenluecke et al., 2011).
P
rotection of coastal
infrastructure
B
iodiversity, tourism Seawalls may provide habitat but these communities have different diversity and structure from those developing on natural
substrates (Jackson et al., 2008); groynes potentially alter beach fauna diversity and community structure (Walker et al., 2008);
c
ontinuing hard protection against sea level rise results in long-term loss of coastal amenities (Gorddard et al., 2012).
A
voidance of risks from sea
level rise via relocation
I
ndigenous communities Relocation can avoid increasing local pressures on communities from sea level rise but raises complex cultural, land rights, legal, and
economic issues, e.g., potential relocation of Torres Strait islander communities (Green, D. et al., 2010a; McNamara et al., 2011).
A
llocating scarce water
resources via market
i
nstruments
R
ural areas, agriculture,
mining
M
arket based instruments such as water trading help allocation of scarce water resources to the highest value uses. The negative
implications of this include potential loss of access to lower value users, which in some areas includes agriculture and drinking
w
ater supplies, with potentially signifi cant social, environmental, and wider economic consequences (Kiem and Austin, 2012).
Increased water security via
augmentation of supply for
u
rban and agricultural systems
Biodiversity, water
demand management
Water storage can buffer urban settlements and agricultural systems against high variability in river fl ows, but altered fl ow
regimes can have signifi cant negative impacts on freshwater ecosystems (Bond et al., 2008; Pittock et al., 2008; Kingsford, 2011).
D
ischarge from desalination plants (e.g. in Perth and Sydney) can lead to substantial local increases in salinity and temperature,
and the accumulation of metals, hydrocarbons, and toxic anti-fouling compounds in receiving waters (Roberts et al., 2010);
i
ncreasing supply can reduce the effectiveness of demand-side measures (Barnett and O’Neill, 2010; Taptiklis, 2011; Box 25-2).
Table 25-6 | Examples of interactions between impacts and adaptation measures in different sectors. In each case, impacts or responses in one sector have the potential to
confl ict (cause negative impacts) or be synergistic (have co-benefi ts) with impacts or responses in another sector, or with another type of response in the same sector.
Primary goal Sector(s) affected Examples of interactions between adaptation and mitigation responses
Adaptation to decreasing
snowfall
Biodiversity, energy
use, water use
Snowmaking in the Australian Alps would require large additional energy and water resources by 2020 of 2500 3300 Ml of water per
month, more than half the average monthly water consumption by Canberra in 2004 2005. Increased snowmaking negatively affects
vegetation, soils, and hydrology of subalpine–alpine areas (Pickering and Buckley, 2010; Morrison and Pickering, 2011; ABS, 2012a).
Air conditioning for heat
stress
Health, energy use Rising temperatures degrade building energy effi ciency (Wang et al., 2010a) and increase energy demand and associated CO
2
emissions
if summer cooling needs are met by increased air conditioning (Stroombergen et al., 2006; Thatcher, 2007; Wang et al., 2010a).
Renewable wind energy
production
Biodiversity Wind-farms can have localized negative effects on bats and birds. However, risk assessment of the potential negative impacts of wind
turbines on threatened bird species in Australia indicated low to negligible impacts on all species modeled (Smales, 2006).
Urban densifi cation Biodiversity, water,
health
Higher urban density to reduce energy consumption from transport and infrastructure can result in loss of permeable surfaces and tree
cover, intensify fl ood risks, and exacerbate discomfort and health impacts of hotter summers (Hamin and Gurran, 2009).
Water supply from
desalination
Energy demand Meeting increasing urban water demand via desalination plants increases energy demand and CO
2
emissions if this demand is met by
increased fossil fuel energy generation (Barnett and O’Neill, 2010; Stamatov and Stamatov, 2010).
Secure food production in
a warming climate
Nitrous oxide and
methane emissions
Net greenhouse gas emissions intensity from dairy systems in southern Australia have been estimated to increase in future in several
locations due to a changing climate and management responses (Cullen and Eckard, 2011; Eckard and Cullen, 2011). A shift toward
perennial C4 grasses would increase methane emissions from grazing ruminants due to lower feed quality, but studies in southwest
Australia suggest this could be more than offset by increased soil carbon storage (Thomas et al., 2012; Bradshaw et al., 2013).
Housing design to reduce
peak energy demand
Energy use,
infrastructure, health
Reducing peak energy demand through building design and demand management reduces vulnerability of electricity networks and
transmission losses during heat waves (Parsons Brinkerhoff, 2009; Nguyen et al., 2010), reduces heat stress during summer, and provides
health benefi ts during winter (Strengers, 2008; Howden-Chapman, 2010; Strengers and Maller, 2011; Ren et al., 2012).
Energy from second-
generation biofuels
Biodiversity, rural
areas, agriculture
New crops such as oil mallees or other eucalypts may provide multiple benefi ts, especially in marginal areas, displacing fossil fuels or
sequestering carbon, generating income for landholders (essential oils, charcoal, bio-char, biofuels), and providing ecosystem services
including reducing erosion (Cocklin and Dibden, 2009; Giltrap et al., 2009; McHenry, 2009).
Reduced emissions from
res
Biodiversity,
livelihoods
Improved management of savannah fi res to reduce the extent of high-intensity late season fi res could substantially reduce emissions as
well as having signifi cant benefi ts for biodiversity and indigenous employment (Russell-Smith et al., 2009; Bradshaw et al., 2013).
Reduce methane
emissions from feral
camels
Biodiversity,
agriculture
Feral camels in Australia are projected to double from 1 to 2 million by 2020. Controlling their numbers to reduce methane emissions
could have signifi cant biodiversity benefi ts (NRMMC, 2010; Bradshaw et al., 2013). Economic benefi ts of reduced grazing competition,
infrastructure damage, and greenhouse gases could outweigh costs of camel reductions (Drucker et al., 2010).
Table 25-7 | Examples of interactions between adaptation and mitigation measures (green rows denote synergies where multiple benefi ts may be realized; yellow rows denote
potential trade-offs and confl icts; blue row gives an example of complex, mixed interactions). The primary goal may be adaptation or mitigation.
1409
Australasia Chapter 25
25
about global climate change impacts, economic effects, and policy
responses.
For New Zealand, there is limited evidence but high agreement that
higher global food prices driven by adverse climate change impacts on
global agriculture and some international climate policies would
increase commodity prices and hence producer returns. Agriculture and
forestry producer returns, for example, are estimated to increase by
14.6% under the A2 scenario by 2070 (Saunders et al., 2010) and real
gross national disposable income by 0.6 to 2.3% under a range of non-
mitigation scenarios (Stroombergen, 2010) relative to baseline projections
in the absence of global climate change.
Some climate policies such as biofuel targets and agricultural mitigation
in other regions would also increase global commodity prices and hence
returns to New Zealand farmers (Saunders et al., 2009; Reisinger et al.,
2012). Depending on global implementation, these could more than offset
projected average domestic climate change impacts on agriculture (Tait
et al., 2008a). In contrast, higher international agricultural commodity
prices appear insufficient to compensate for the more severe effects of
climate change on agriculture in Australia (see Section 25.7.2; Gunasekera
et al., 2007; Garnaut, 2008).
Climate change could affect international tourism to Australasia through
international destination and activity preferences (Kulendran and Dwyer,
2010; Rosselló-Nadal et al., 2011; Scott et al., 2012), climate policies,
and oil prices (Mayor and Tol, 2007; Becken, 2011; Schiff and Becken,
2011). These potentially significant effects remain poorly quantified,
however, and are not well integrated into local vulnerability studies
(Hopkins et al., 2012).
Box 25-10 | Land-based Interactions among Climate, Energy, Water, and Biodiversity
Climate, water, biodiversity, food, and energy production and use are intertwined through complex feedbacks and trade-offs (see also
Box CC-WE). This could make alternative uses of natural resources within rural landscapes increasingly contested, yet decision support
tools to manage competing objectives are limited (PMSEIC, 2010b).
Various policies in Australasia support increased biofuel production and biological carbon sequestration via, for example, mandatory
renewable energy targets and incentives to increase carbon storage. Impacts of increased biological sequestration activities on
biodiversity depend on their implementation. Benefits arise from reduced erosion, additional habitat, and enhanced ecosystem
connectivity, while risks or lost opportunities are associated with large-scale monocultures especially if replacing more diverse
landscapes (Brockerhoff et al., 2008; Giltrap et al., 2009; Steffen et al., 2009; Todd et al., 2009; Bradshaw et al., 2013).
Photosynthesis transfers water to the atmosphere, so increased sequestration is projected to reduce catchment yields particularly in
southern Australia and affect water quality negatively (CSIRO, 2008; Schrobback et al., 2011; Bradshaw et al., 2013). Accounting for
this water use in water allocations for sequestration activities would increase their cost and limit the potential of sequestration-driven
land use change (Polglase et al., 2011; Stewart et al., 2011). Large-scale land-cover changes also affect local and regional climates
and soil moisture through changing albedo, evaporation, plant transpiration, and surface roughness (McAlpine et al., 2009;
Kirschbaum et al., 2011b), but these feedbacks have rarely been included in analyses of changing water demands and availability.
Biological carbon sequestration in New Zealand is less water-challenged than in Australia, except where catchments are projected to
become drier and/or are already completely allocated (MfE, 2007a; Rutledge et al., 2011), and would mostly improve water quality
through reduced erosion (Giltrap et al., 2009). Policies to protect water quality by limiting nitrogen discharge from agriculture have
reduced livestock production and greenhouse gas emissions in the Lake Taupo and Rotorua catchments and supported land-use
change toward sequestration (OECD, 2013b).
Trade-offs between biofuel and food production and ecosystem services depend strongly on the type of sequestration activity and
their management relies on the use of consistent principles to evaluate externalities and benefits of alternative land uses (PMSEIC,
2010b). First-generation biofuels have been modeled in Australia as directly competing with agricultural production (Bryan et al.,
2010). In contrast, production of woody biofuels in New Zealand is projected to occur on marginal land, not where the most intense
agriculture occurs (Todd et al., 2009). Falling costs and increasing efficiency of solar energy may limit future biofuel demand, given
the limited efficiency of plants in converting solar energy into usable fuel (e.g., Reijnders and Huijbregts, 2007).
1410
Chapter 25 Australasia
25
C
limate change has the potential to change migration flows within
Australasia, particularly because of coastal changes (e.g., from the Torres
Straits islands to mainland Australia), although reliable estimates of
such movements do not yet exist (see Section 12.4; Green, D. et al.,
2010a; McNamara et al., 2011; Hugo, 2012). Migration within countries,
and from New Zealand to Australia, is largely economically driven and
sustained by transnational networks, though the perceived more attractive
current climate in Australia is reportedly a factor in migration from New
Zealand (Goss and Lindquist, 2000; Green, A.E. et al., 2008; Poot, 2009).
The impacts of climate change in the Pacific may contribute to an increase
in the number of people seeking to move to nearby countries (Bedford
and Bedford, 2010; Hugo, 2010; McAdam, 2010; Farbotko and Lazrus,
2012; Bedford and Campbell, 2013) and affect political stability and
geopolitical rivalry within the Asia-Pacific region, although there is no
clear evidence of this to date and causal theories are scarce (Dupont,
2008; Pearman, 2009; see Sections 12.4-5). Increasing climate-driven
disasters, disease, and border control will stimulate operations other
than war for Australasia’s armed forces; integration of security into
adaptation and development assistance for Pacific island countries can
therefore play a key role in moderating the influence of climate change
on forced migration and conflict (robust evidence, high agreement;
Dupont and Pearman, 2006; Bergin and Townsend, 2007; Dupont, 2008;
Sinclair, 2008; Barnett, 2009; Rolfe, 2009).
25.10. Synthesis and Regional Key Risks
25.10.1. Economy-wide Impacts and
the Potential of Mitigation to Reduce Risks
Globally effective mitigation could reduce or delay some of the risks
associated with climate change and make adaptation more feasible
beyond about 2050, when projected climates begin to diverge
substantially between mitigation and non-mitigation scenarios (see also
Section 19.7). Literature quantifying these benefits for Australasia has
increased since the AR4 but remains very sparse. Economy-wide net
costs for Australia are modeled to be substantially greater in 2100 under
unmitigated climate change (A1FI; GNP loss 7.6%) than under globally
effective mitigation (GNP loss less than 2% for stabilization at 450 or
550 ppm CO
2
-eq, including costs of mitigation and residual impacts;
Garnaut, 2008). These estimates, however, are highly uncertain and
depend strongly on valuation of non-market impacts, treatment of
potentially catastrophic outcomes, and assumptions about adaptation,
global changes, and flow-on effects for Australia and effectiveness and
i
mplementation of global mitigation efforts (Garnaut, 2008). No
estimates of climate change costs across the entire economy exist for
New Zealand.
The benefits of mitigation in terms of reduced risks have been quantified
for some individual sectors in Australia, for example, for irrigated
agriculture in the Murray-Darling Basin (Quiggin et al., 2008, 2010;
Valenzuela and Anderson, 2011; Scealy et al., 2012) and for net health
outcomes (Bambrick et al., 2008). Although quantitative estimates from
individual studies are highly assumption-dependent, multiple lines of
evidence (see Sections 25.7-8) give very high confidence that globally
effective mitigation would significantly reduce many long-term risks
from climate change to Australia. Benefits differ, however, between
States for some issues, for example, heat- and cold-related mortality
(Bambrick et al., 2008). Few studies consider mitigation benefits explicitly
for New Zealand, but scenario-based studies give high confidence that,
if global emissions were reduced from a high (A2) to a medium-
low (B1) emissions scenario, this would markedly lower the projected
increase in flood risks (Ballinger et al., 2011; McMillan et al., 2012) and
reduce risks to livestock production in the most drought-prone regions
(Tait et al., 2008a; Clark et al., 2011). Mitigation would also reduce the
projected benefits to production forestry, however, though amounts
depend on the response to CO
2
fertilization (Kirschbaum et al., 2011a;
see also Section 25.7.1).
25.10.2. Regional Key Risks as a Function
of Mitigation and Adaptation
The Australia/New Zealand chapter of the AR4 (Hennessy et al., 2007)
concluded with an assessment of aggregated vulnerability for a range
of sectors as a function of global average temperature. Building on
recent additional insights, Table 25-8 shows eight key risks within those
sectors that can be identified with high confidence for the 21st century,
based on the multiple lines of evidence presented in the preceding
sections and selected using the framework for identifying key risks set
out in Chapter 19 (see also Box CC-KR). This combines consideration of
biophysical impacts, their likelihood, timing, and persistence, with
vulnerability of the affected system, based on exposure, magnitude of
harm, significance of the system, and its ability to cope with or adapt
to projected biophysical changes. These key risks differ in the extent to
which they can be managed through adaptation and mitigation and
their evolution over time, and some are more likely than others, but all
warrant attention from a risk-management perspective.
Table 25-8 | Key regional risks from climate change and the potential for reducing risk through mitigation and adaptation. Key risks are identified based on assessment
of the literature and expert judgments by chapter authors, with evaluation of evidence and agreement in the supporting chapter sections. Each key risk is characterized on
a scale from very low to very high and presented in three timeframes: the present, near-term (2030–2040), and long-term (2080–2100). For the near-term era of
committed climate change (here, for 2030–2040), projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. For the
longer-term era of climate options (here, for 2080–2100), risk levels are presented for global mean temperature increase of 2°C and 4°C above preindustrial levels. For
each timeframe, risk levels are estimated for a continuation of current adaptation and for a hypothetical highly adapted state. Relevant climate variables are indicated by
icons. For a given key risk, change in risk level through time and across magnitudes of climate change is illustrated, but because 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, sectors, or regions.
1411
Australasia Chapter 25
25
Present
Long term
(20802100)
2°C
4°C
V
ery
l
ow
V
ery
h
igh
Medium
Present
Long term
(20802100)
2°C
4°C
Very
low
Very
high
Medium
Present
Long term
(20802100)
2°C
4°C
Very
low
Very
high
Medium
Near term
(2030–2040)
Present
Long term
(20802100)
2°C
4°C
Very
low
Very
high
Medium
Present
Long-term
(2080-2100)
2°C
4°C
Very
low
Very
high
Medium
1.5°C
Near-term
(2030-2040)
1.5°C
Near term
(2030–2040)
1.5°C
Near term
(2030–2040)
1.5°C
Near term
(2030–2040)
1.5°C
I
mpacts can be delayed but now appear very difficult to avoid entirely, even with combined globally effective mitigation and planned adaptation
Impacts have the potential to be severe but can be reduced substantially by globally effective mitigation combined with adaptation
Impacts whose severity depends on changes in climate variables that span a particularly large range; the most severe end would present major challenges
Near term
(2030–2040)
Present
Long term
(20802100)
2°C
4°C
Very
low
Very
high
Medium
Present
Long-term
(2080-2100)
2°C
4°C
Very
low
Very
high
Medium
1.5°C
Near-term
(2030-2040)
1.5°C
Loss of montane ecosystems and some native
s
pecies in Australia (high confidence)
[25.6.1]
Direct adaptation options are limited, but reducing other stresses such as
p
ests and diseases, predator control and enhancing connectivity of habitats
provides immediate co-benefits; need to consider facilitating migration and
assisted colonisation.
Constraints on water resources in southern
Australia (high confidence)
[25.5.1, Boxes 25-2, 25-9]
Water resources already struggling to meet unrestrained demand in many
locations and exacerbated by projected population growth; effective
adaptation relies on combination of demand and supply mechanisms.
Increased morbidity, mortality and
infrastructure damages during heat waves in
Australia (high confidence)
[25.7.4, 25.8.1]
Vulnerability is exacerbated by population growth and aging; transport and
power infrastructure already severely stressed during heat waves in many
regions, with significant financial costs from future upgrades.
Significant reduction in agriculture
production in the Murray-Darling Basin
and far south-eastern and
south-western Australia
(high confidence)
Wet end of scenario
Dry end of scenario
Increasing risks to coastal infrastructure
and low-lying ecosystems in Australia and
New Zealand, with widespread damages
toward the upper end of projected sea
level rise ranges (high confidence)
[25.6, 25.10, Box 25-1]
Moderate sea level rise High end sea level rise
Immediate co-benefits from improved
management of over-allocated water
resources and balancing competing
demands, but the extreme dry end would
threaten agricultural production as well as
ecosystems and some rural communities.
(25.2, 25.5.2, Figure 25-4)
(AR5 WGI 13.5; Box 25-2)
[25.2, 25.6.1, 25.7.2, Table 25-1,
Boxes 25-2, 25-5]
Present
Long term
(20802100)
2°C
4°C
Very
low
Very
high
Medium
Near term
(2030–2040)
1.5°C
Wild fire damages to ecosystems and
settlements and risks to human life in southern
Australia and many parts of New Zealand
(high confidence)
[Table 25-1, Box 25-6]
Part of integrated landscape management; trade-offs between different
management objectives and settlement patterns and goals (biodiversity
versus protection of human life and property).
Adaptation deficit in some locations to
current coastal erosion and flood risk.
Successive building and protection cycles
constrain flexible responses. Effective
adaptation includes land-use controls and
ultimately relocation as well as protection
and accommodation.
Near term
(20302040)
Present
Long term
(20802100)
2
°C
4°C
Very
low
Very
high
Medium
Significant change in community composition
a
nd structure of coral reef systems in Australia
(high confidence)
[
25.6.2, 30.5, Boxes CC-CR, CC-OA]
Ability of corals to adapt naturally appears limited and insufficient to
o
ffset the detrimental effects of rising temperatures and
acidification.Other options are mostly limited to reducing other stresses
(water quality, tourism, fishing) and early warning systems; direct
i
nterventions such as assisted colonization and shading have been
proposed but remain untested at scale.
Near term
(2030–2040)
Present
L
ong term
(2080–2100)
2
°C
4°C
Very
low
Very
high
Medium
Increased frequency and intensity of flood
d
amage to infrastructure and settlements in
Australia and New Zealand (high confidence)
[Table 25-1, Boxes 25-8, 25-9]
Significant adaptation deficit in some regions to current flood risk. Effective
adaptation includes land-use controls and relocation as well as protection
a
nd accommodation of increased risk to ensure flexibility.
1.5°C
C
OO
Key risk Adaptation issues & prospects
Climatic
drivers
Risk & potential for
adaptation
Timeframe
D
amaging
c
yclone
O
cean
a
cidification
C
OO
Climate-related drivers of impacts
W
arming
t
rend
E
xtreme
p
recipitation
E
xtreme
t
emperature
Level of risk & potential for adaptation
P
otential for additional adaptation
t
o reduce risk
Risk level with
c
urrent adaptation
Risk level with
h
igh adaptation
D
rying
t
rend
S
now
c
over
S
ea level
r
ise
1.5°C
1412
Chapter 25 Australasia
25
O
ne set of key risks comprises damages to natural ecosystems (significant
change in community structure of coral reefs and loss of some montane
ecosystems) that can be moderated by globally effective mitigation but
to which some damage now seems inevitable. For some species and
ecosystems, climatically constrained ecological niches, fragmented
habitats, and limited adaptive movement collectively present hard limits
to adaptation to further climate change (high confidence). A second set
of key risks (increase in flood risk, water scarcity, heat waves, and
wildfire) comprises damages that could be severe but can be reduced
substantially by globally effective mitigation combined with adaptation,
with the need for transformational adaptation increasing with the rate
and amount of climate change. A third set of key risks (coastal damages
from sea level rise, and loss of agriculture production from severe drying)
comprises potential impacts whose scale remains highly uncertain
within the 21st century, even for a given global temperature change,
and where alternative scenarios materially affect levels of concern,
adaptation needs, and strategies. Even though scenarios of severe drying
(see Section 25.5.2) or rapid sea level rise approaching 1 m or more by
2100 (see Box 25-2; WGI AR5 Section 13.5) have low or currently
unknown probabilities, the associated impacts would so severely
challenge adaptive capacity, including transformational changes, that
they constitute important risks.
A first comparative assessment for Australia of exposure and damages
from different hazards up to 2100 indicates that river flooding will
continue to be the most costly source of direct damages to infrastructure,
even though the largest value of assets is exposed to bush fire. Exposure
to and damages from coastal inundation are currently smaller, but
would rise most rapidly beyond mid-century if sea level rise exceeds
0.5 m (Baynes et al., 2012).
An emerging risk is the compounding of extreme events, none of which
would constitute a key risk in its own right, but that collectively and
cumulatively across space and time could stretch emergency response
and recovery capacity and hamper regional economic development,
including through impacts on insurance markets or multiple concurrent
needs for major infrastructure upgrades (NDIR, 2011; Phelan, 2011;
Baynes et al., 2012; Booth and Williams, 2012; Karoly and Boulter, 2013).
Efforts are underway to better understand the potential importance of
cumulative impacts and responses, including the challenges arising from
impacts and responses across different levels of government (Leonard,
M. et al., 2010; CSIRO, 2011), but evidence is as yet too limited to
identify this as a key risk consistent with the definitions adopted in this
report (see Chapter 19).
Climate change is projected to bring benefits to some sectors and parts
of Australasia, at least under limited warming scenarios associated with
globally effective mitigation (high confidence). Examples include an
extended growing season for agriculture and forestry in cooler parts of
New Zealand and Tasmania, reduced winter mortality (low confidence),
and reduced winter energy demand in most of New Zealand and southern
States of Australia, and increased winter hydropower potential in New
Zealand’s South Island (Sections 25.7.1-2, 25.7.4, 25.8.1).
The literature supporting this assessment of key risks is uneven among
sectors and between Australia and New Zealand; for the latter, conclusions
in many sectors are based on limited studies that often use a narrow
s
et of assumptions, models, and data and that, accordingly, have not
explored the full range of potential outcomes.
25.10.3. Challenges to Adaptation in Managing Key Risks,
and Limits to Adaptation
Two key and related challenges for regional adaptation are apparent:
to identify when and where adaptation may imply transformational
rather than incremental changes; and, where specific interventions are
needed to overcome adaptation constraints, in particular to support
transformational responses that require coordination across different
spheres of governance and decision making (Productivity Commission,
2012; Palutikof et al., 2013). The magnitude of climate change, especially
under scenarios of limited mitigation, and constraints to adaptation
suggest that incremental and autonomous responses will not deliver
the full range of available adaptation options nor ensure the continued
function of natural and human systems if some key risks are realized
(high confidence; see also Section 25.4).
Most incremental adaptation measures in natural ecosystems focus on
reducing other non-climate stresses but, even with scaled-up efforts,
conserving the current state and composition of the ecosystems most
at risk appears increasingly infeasible (Sections 25.6.1-2). Maintenance
of key ecosystem functions and services requires a radical reassessment
of conservation values and practices related to assisted colonization
and the values placed on “introduced” species (Steffen et al., 2009).
Divergent views regarding intrinsic and service values of species and
ecosystems imply the need for a proactive discussion to enable effective
decision making and resource allocation.
In human systems, incremental adjustments of current risk management
tools, planning approaches, and early warning systems for floods, fire,
drought, water resources, and coastal hazards can increase resilience
to climate variability and change, especially in the near term (IPCC, 2012;
Productivity Commission, 2012; Dovers, 2013). A purely incremental
approach, however, which generally aims to preserve current management
objectives, governance, and institutional arrangements, can make later
transformational changes increasingly difficult and costly (medium
evidence, high agreement; e.g., Howden et al., 2010; Park et al., 2012;
McDonald, 2013; Stafford-Smith, 2013). Examples of transformational
changes include: shifting emphasis from protection to accommodation
or avoidance of flood risk, including managed retreat from eroding
coasts; the translocation of industries in response to increasing drought,
flood, and fire risks or water scarcity; and the associated transformation
of the economic and social base and governance of some rural
communities (Boxes 25-1, 25-2, 25-5 to 25-9; Nelson et al., 2010;
Linnenluecke et al., 2011; Kiem and Austin, 2012; O’Neill and Handmer,
2012; McDonald, 2013; Palutikof et al., 2013).
Consideration of transformational adaptation becomes critical where
long life- or lead-times are involved, and where high up-front costs or
multiple interdependent actors create constraints that require coordinated
and proactive interventions (Stafford-Smith et al., 2011; Productivity
Commission, 2012; Palutikof et al., 2013). Deferring such adaptation
decisions because of uncertainty about the future will not necessarily
minimize costs or ensure adequate flexibility for future responses,
1413
Australasia Chapter 25
25
although up-front investment and opportunity costs of adaptation
can present powerful arguments for delayed or staged responses
(Stewart and Wang, 2011; Gorddard et al., 2012; Productivity Commission,
2012; McDonald, 2013). Whether transformational responses are seen
as success or failure of adaptation depends on the extent to which
actors accept a change in, or wish to maintain, current activities and
management objectives, and the degree to which the values and
institutions underpinning the transformation are shared or contested
across stakeholders (Park et al., 2012; Stafford-Smith, 2013). These views
will differ not only between communities and industries but also from
person to person depending on their individual value systems, perceptions
of and attitude to risk, and ability to capitalize on opportunities (see
also Section 25.4.3).
25.11. Filling Knowledge Gaps to Improve
Management of Climate Risks
The wide range of projected rainfall changes (averages and extremes)
and their hydrological amplification are key uncertainties affecting the
scale and urgency of adaptation in agriculture, forestry, water resources,
some ecosystems, and wildfire and flood risks. For ecosystems, agriculture,
and forestry, these uncertainties are compounded by limited knowledge
of responses of vegetation to elevated CO
2
, changes in ocean pH, and
interactions with changing climatic conditions. The uncertainties in future
impacts are most critical for decisions with long lifetimes, such as capital
infrastructure investment or large-scale changes in land and water use.
Uncertainties about the rate of sea level rise, and changes in storm
paths and intensity, add to challenges for infrastructure design. The use
of multi-model means and a narrow set of emissions scenarios in many
past studies implies that the full set of climate-related risks and
management options remains incompletely explored.
Understanding of ecological and physiological thresholds that, once
exceeded, would result in rapid changes in species, ecosystems, and their
services is still very limited. The literature is noticeably sparse in New
Zealand and for arid Australia. These knowledge gaps are compounded
by limited information about the effect of global climate change on
patterns of natural climate variability, such as ENSO. Better understanding
the effect of evolving natural climate variability and long-term trends,
along with rising CO
2
concentrations, on pests, invasive species, and
native and managed ecosystems could support more robust ecosystem-
based adaptation strategies.
Vulnerability of human and managed systems depends critically on
future socioeconomic characteristics. Research into psychological,
economic, social, and cultural dimensions of vulnerability, adaptive
capacity, and underpinning values remains limited and poorly integrated
with biophysical studies. This limits the level of confidence in conclusions
regarding future vulnerabilities and the feasibility and effectiveness of
adaptation strategies.
These multiple, persistent, and structural uncertainties imply that, in
most cases, adaptation requires an iterative risk management process.
Though decision-support frameworks are being developed, it remains
unclear to what extent existing governance and institutional arrangements
will be able to support more transformational responses, particularly
where competing public and private interests and particularly vulnerable
groups are involved. The enabling or constraining influences on
adaptation from interactions among market forces, institutions,
governance, policy, and regulatory environments have only recently
begun to attract research attention, mostly in Australia.
Climate change impacts, adaptation, and mitigation responses in other
world regions will affect Australasia, but our understanding of this
Frequently Asked Questions
FAQ 25.2 | What are the key risks from climate change to Australia and New Zealand?
Our assessment identifies eight key regional risks from climate change. Some impacts, especially on ecosystems,
a
re by now difficult to avoid entirely. Coral reef systems have a limited ability to adapt naturally to further warming
and an increasingly acidic ocean. Similarly, the habitat for some mountain or high elevation ecosystems and their
associated species is shrinking inexorably with rising temperatures. This implies substantial impacts and some losses
e
ven under scenarios of limited warming. Other risks, however, can be reduced substantially by adaptation, combined
with globally effective mitigation. These include potential flood damages from more extreme rainfall in most parts
of Australia and New Zealand; constraints on water resources from reducing rainfall in southern Australia; increased
h
ealth risks and infrastructure damages from heat waves in Australia; and increased economic losses, risks to human
life, and ecosystem damage from wildfires in southern Australia and many parts of New Zealand. A third set of
risks is particularly challenging to manage robustly because the severity of potential impacts varies widely across
the range of climate projections, even for a given temperature increase. These concern damages to coastal
infrastructure and low-lying ecosystems from continuing sea level rise, where damages would be widespread if
sea level turns out to be at the upper end of current scenarios; and threats to agricultural production in both far
southeastern and far southwestern Australia, which would affect ecosystems and rural communities severely at the
dry end of projected rainfall changes. Even though some of these key risks are more likely to materialize than others,
and they differ in the extent that they can be managed by adaptation and mitigation, they all warrant attention
from a risk management perspective, given their potential major consequences for the region.
1414
Chapter 25 Australasia
25
r
emains very limited. Existing studies suggest that transboundary
effects, mediated mostly via trade but potentially also migration, can
be of similar if not larger scale than direct domestic impacts of climate
change for economically important sectors such as agriculture and
tourism. However, scenarios used in such studies tend to be highly
simplified. Effective management of risks and opportunities in these
sectors would benefit from better integration of relevant global scenarios
of climatic and socioeconomic changes into studies of local vulnerability
and adaptation options.
References
ABARES, 2011: Australian Mineral Statistics 2011, March Quarter 2011. ABARES
Project 43133, Australian Bureau of Agricultural and Resource Economics and
Sciences (ABARES), Canberra, ACT, Australia, 37 pp.
ABARES, 2012: Australia’s Forests at a Glance 2012. Australian Bureau of Agricultural
and Resource Economics and Sciences (ABARES), Canberra, Australia, ACT, 83 pp.
Abbs, D.J., 2012: The Impact of Climate Change on the Climatology of Tropical Cyclones
in the Australian Region. CSIRO Climate Adaptation Flagship Working Paper
No. 11, Commonwealth Scientific and Industrial Research Organization (CSIRO),
Aspendale, VIC, Australia, 17 pp.
Abbs, D.J. and A. Rafter, 2009: Impact of Climate Variability and Climate Change on
Rainfall Extremes in Western Sydney and Surrounding Areas: Component 4
Dynamical Downscaling. Report to the Sydney Metro Catchment Management
Authority and Partners, Commonwealth Scientific and Industrial Research
Organization (CSIRO), Aspendale, VIC, Australia, 83 pp.
Abdo, D.A., D.A. Bellchambers, L.M. Evans, and S.N. Evans, 2012: Turning up the heat:
increasing temperature and coral bleaching at the high latitude coral reefs of
the Houtman Abrolhos Islands. PLoS ONE, 7(8), e43878, doi:10.1371/journal.
pone.0043878.
Abel, N., R. Gorddard, B. Harman, A. Leitch, J. Langridge, A. Ryan, and S. Heyenga,
2011: Sea level rise, coastal development and planned retreat: analytical
framework, governance principles and an Australian case study. Environmental
Science & Policy, 14(3), 279-288.
ABS, 2008: Population Projections Australia, 2006 to 2101. Catalogue No. 3222
Australian Bureau of Statistics (ABS), Canberra, ACT, Australia, 96 pp.
ABS, 2009: Experimental Estimates and Projections, Aboriginal and Torres Strait
Islander Australians. Catalogue 3238.0, Australian Bureau of Statistics (ABS),
Canberra, ACT, Australia, 76 pp.
ABS, 2010a: Tourism Satellite Account 2009-10. Catalogue No. 5249.0, Australian
Bureau of Statistics (ABS), Canberra, ACT, Australia, 50 pp.
ABS, 2010b: Population Characteristics, Aboriginal and Tores Trait Islander Australians.
Catalogue No. 4713.0, Australian Bureau of Statistics (ABS), Canberra, ACT,
Australia, 200 pp.
ABS, 2010c: The Health and Welfare of Australia’s Aboriginal and Torres Strait Islander
Peoples, 2010. Catalogue 4704.0, Australian Bureau of Statistics (ABS),
Canberra, ACT, Australia, 107 pp.
ABS, 2012a: Water Account Australia 2009-10. Catalogue 4610.0, Australian Bureau
of Statistics (ABS), Canberra, ACT, Australia, 93 pp.
ABS, 2012b: Gross Value of Irrigated Agricultural Production, 2010-11. Catalogue No.
4610.0.55.008, Australian Bureau of Statistics (ABS), Canberra, ACT, Australia,
www.abs.gov.au/AUSSTATS/abs@.nsf/DetailsPage/4610.0.55.0082010-11.
ABS, 2012c: Yearbook Australia 2012. Catalogue No. 1301.0, Australian Bureau of
Statistics (ABS), Canberra, ACT, Australia, 885 pp.
ABS, 2012d: Australian National Accounts: National Income, Expenditure and
Product, June Quarter 2012: Table 5206.0. Australian Bureau of Statistics (ABS),
Canberra, ACT, Australia, 81 pp.
ABS, 2013: Australian Demographic Statistics, December Quarter 2012: Table 3101.0.
Australian Bureau of Statistics (ABS), Canberra, ACT, Australia, 104 pp.
Ackerley, D., R. Bell, B. Mullan, and H. McMillan, 2013: Estimation of regional
departures from global-average sea-level rise around New Zealand from
AOGCM simulations. Weather and Climate, 33, 2-22.
ACT Government, 2012: AP2: A New Climate Change Strategy and Action Plan for
the Australian Capital Territory. Environment and Sustainable Development
Directorate, Australian Capital Territory (ACT) Government, Canberra, ACT,
Australia, 94 pp.
Adams, M. and P. Attiwill, 2011: Burning Issues: Sustainability and Management of
Australia’s Southern Forests. Commonwealth Scientific and Industrial Research
Organization (CSIRO) Publishing, Collingwood, VIC, Australia, 144 pp.
Adams-Hosking, C., H.S. Grantham, J.R. Rhodes, C. McAlpine, and P.T. Moss, 2011:
Modelling climate-change-induced shifts in the distribution of the koala.
Wildlife Research, 38(2), 122-130.
Agho, K., G. Stevens, M. Taylor, M. Barr, and B. Raphael, 2010: Population risk perceptions
of global warming in Australia. Environmental Research, 110(8), 756-763.
AGO, 2006: Climate Change Impacts and Risk Management: A Guide for Business and
Government. Australian Greenhouse Office (AGO), Department of Environment
and Heritage, Canberra, ACT, Australia, 73 pp.
Akompad, D.A., P. Bi, S. Williams, J. Grant, I.A. Walker, and M. Augoustinous, 2013:
Heat waves and climate change: applying the health belief model to identify
predictors of risk preception and adapative behaviours in Adelaide, Australia.
International Journal of Environmental Reserach and Public Health, 10, 2164-2184.
Alexander, K.S., A. Ryan, and T.G. Measham, 2012: Managed retreat of coastal
communities: understanding responses to projected sea level rise. Journal of
Environmental Planning and Management, 55(4), 409-433.
Alexander, L.V. and J.M. Arblaster, 2009: Assessing trends in observed and modelled
climate extremes over Australia in relation to future projections. International
Journal of Climatology, 29(3), 417-435.
Alexander, L.V., P. Hope, D. Collins, B. Trewin, A. Lynch, and N. Nicholls, 2007:
Trends in Australia’s climate means and extremes: a global context. Australian
Meteorological Magazine, 56(1), 1-18.
Alexander, L.V., X.L. Wang, H. Wan, and B. Trewin, 2011: Significant decline in
storminess over southeast Australia since the late 19
t
h
century. Australian
Meteorological and Oceanographic Journal, 61, 23-30.
Allen, C.D., A.K. Macalady, H. Chenchouni, D. Bachelet, N. McDowell, M. Vennetier, T.
Kitzberger, A. Rigling, D.D. Breshears, E.H. Hogg, P. Gonzalez, R. Fensham, Z.
Zhang, J. Castro, N. Demidova, J.H. Lim, G. Allard, S.W. Running, A. Semerci, and
N. Cobb, 2010: A global overview of drought and heat-induced tree mortality
reveals emerging climate change risks for forests. Forest Ecology and
Management, 259(4), 660-684.
Allen, J.T. and D.J. Karoly, 2013: A climatology of Australian severe thunderstorm
environments 1979-2011: inter-annual variability and ENSO influence.
International Journal of Climatology (in press), doi:10.1002/joc.3667.
Alston, M., 2007: Gender and climate change: variable adaptations of women and
men. Just Policy, 46, 29-35.
Alston, M., 2010: Gender and climate change in Australia. Journal of Sociology,
47(1), 53-70.
Alston, M., 2012: Rural male suicide in Australia. Social Science & Medicine, 74(4),
512-522.
Altman, J., G. Buchanan, and L. Larsen, 2007: The Environmental Significance of the
Indigenous Estate: Natural Resource Management as Economic Development
in Remote Australia. CAEPR Discussion Paper No. 286/2007, Centre for Aboriginal
Economic Policy Research (CAEPR), The Australian National University (ANU),
Canberra, ACT, Australia, 65 pp.
Anderson, B., W. Lawson, and I. Owens, 2008: Response of Franz Josef Glacier Ka
Roimata o Hine Hukatere to climate change. Global and Planetary Change,
63(1), 23-30.
Anderson, I., S. Crengle, M. Leialoha Kamaka, T.-H. Chen, N. Palafox, and L. Jackson-
Pulver, 2006: Indigenous health in Australia, New Zealand, and the Pacific. The
Lancet, 367(9524), 1775-1785.
Anthony, K.R.N., D.I. Kline, G. Diaz-Pulido, S. Dove, and O. Hoegh-Guldberg, 2008:
Ocean acidification causes bleaching and productivity loss in coral reef builders.
Proceedings of the National Academy of Sciences of the United States of
America, 105(45), 17442-17446.
ANU, 2009: Implications of Climate Change for Australia’s World Heritage Properties:
A Preliminary Assessment. Report to the Department of Climate Change (DCC)
and the Department of the Environment, Water, Heritage and the Arts (DEWHA)
by the Fenner School of Environment and Society, The Australian National
University (ANU), Canberra, ACT, Australia, 207 pp.
Arkema, K.K., G. Guannel, G. Verutes, S.A. Wood, A. Guerry, M. Ruckelshaus, P.
Kareiva, M. Lacayo, and J.M. Silver, 2013: Coastal habitats shield people and
property from sea-level rise and storms. Nature Climate Change 3, 913-918.
Ashworth, P., T. Jeanneret, J. Gardner, and H. Shaw, 2011: Communication and
Climate Change: What the Australian Public Thinks. Report EP112769,
Commonwealth Scientific and Industrial Research Organization (CSIRO),
Canberra, ACT, Australia, 68 pp.
1415
Australasia Chapter 25
25
Asseng, S., F. Ewert, C. Rosenzweig, J.W. Jones, J.L. Hatfield, A.C. Ruane, K.J. Boote,
P.J. Thorburn, R.P. Rotter, D. Cammarano, N. Brisson, B. Basso, P. Martre, P.K.
Aggarwal, C. Angulo, P. Bertuzzi, C. Biernath, A.J. Challinor, J. Doltra, S. Gayler, R.
Goldberg, R. Grant, L. Heng, J. Hooker, L.A. Hunt, J. Ingwersen, R.C. Izaurralde,
K.C. Kersebaum, C. Muller, S. Naresh Kumar, C. Nendel, G. O’Leary, J.E. Olesen,
T.M. Osborne, T. Palosuo, E. Priesack, D. Ripoche, M.A. Semenov, I. Shcherbak, P.
Steduto, C. Stockle, P. Stratonovitch, T. Streck, I. Supit, F. Tao, M. Travasso, K. Waha,
D. Wallach, J.W. White, J.R. Williams, and J. Wolf, 2013: Uncertainty in simulating
wheat yields under climate change. Nature Climate Change, 3(9), 827-832.
Åström, C., J. Rocklöv, S. Hales, A. Béguin, V. Louis, and R. Sauerborn, 2012: Potential
distribution of dengue fever under scenarios of climate change and economic
development. EcoHealth, 9(4), 448-454.
ATSE, 2008: Assessment of Impacts of Climate Change on Australia’s Physical
Infrastructure. The Australian Academy of Technological Sciences and Engineering
(ATSE), Parkville, VIC, Australia, 55 pp.
ATSISJC, 2009: Native Title Report 2008. Aboriginal and Torres Strait Islander Social
Justice Commissioner (ATSISJC), Australian Human Rights Commission, Sydney,
NSW, Australia, 407 pp.
Auckland Council, 2012: The Auckland Plan. Auckland Council, Auckland, New
Zealand, 372 pp.
August, S.M. and B.J. Hicks, 2008: Water temperature and upstream migration of
glass eels in New Zealand: implications of climate change. Environmental
Biology of Fishes, 81(2), 195-205.
Australian Government, 2012: State of Australian Cities 2012. Department of
Infrastructure and Transport, Australian Government, Canberra, ACT, Australia,
365 pp.
Australian Treasury, 2010: The 2010 Intergenerational Report. Australia to 2050:
Future Challenges. Australian Treasury, Canberra, ACT, Australia, 164 pp.
Bacon, W., 2011: A Sceptical Climate. Media Coverage of Climate Change in Australia.
Part 1 – Climate Change Policy. Australian Centre for Independent Journalism
(ACIJ), University of Technology, Sydney, NSW, Australia, 70 pp.
Baisden, W.T., E.D. Keller, L. Timar, D. Smeaton, A. Clark, A. Ausseil, W. Power, and W.
Zhang, 2010: New Zealands Pasture Production in 2020 and 2050. GNS Sciences
Consultancy Report 2010/154 prepared by the Institute of Geological and
Nuclear Sciences Ltd. (GNS Sciences) for the Ministry of Agriculture and
Forestry, GNS Sciences, Lower Hutt, New Zealand, 62 pp.
Ballinger, J., B. Jackson, A. Reisinger, and K. Stokes, 2011: The Potential Effects of
Climate Change on Flood Frequency in the Hutt River. School of Geography,
Environment and Earth Sciences, Victoria University of Wellington, Wellington,
New Zealand, 40 pp.
Balston, J.M., 2012: Guidelines for Developing a Climate Change Adaptation Plan
and Undertaking an Integrated Climate Change Vulnerability Assessment. Local
Government Association (LGA) of South Australia, Adelaide, SA, Australia, 61 pp.
Balston, J.M., J. Kellett, G. Wells, S. Li, A. Gray, and I. Iankov, 2012: Quantifying the
Costs of Climate Change on Local Government Assets. Local Government
Association (LGA) of South Australia and National Climate Change Adaptation
Research Facility (NCCARF), NCCARF, Griffith University Gold Coast Campus,
QLD, Australia, 219 pp.
Bambrick, H.J., K. Dear, R. Woodruff, I. Hanigan, and A.M. Michael, 2008: The Impacts
of Climate Change on Three Health Outcomes: Temperature-Related Mortality
and Hospitalisations, Salmonellosis and other Bacterial Gastroenteritis, and
Population at Risk from Dengue. Paper prepared for Garnaut Climate Change
Review, Garnaut Climate Change Review, Canberra, ACT, Australia, 45 pp.
Banfai, D.S. and D. Bowman, 2007: Drivers of rain-forest boundary dynamics in
Kakadu National Park, northern Australia: a field assessment. Journal of Tropical
Ecology, 23, 73-86.
Banks, S.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.
Barnett, J., 2009: Climate change and human security in the Pacific Islands: the
potential for and limits to adaptation. In: Climate Change and Security. Planning
for the Future [Boston, J., P. Nel, and M. Righarts (eds.)]. Institute of Policy Studies,
Victoria University of Wellington, Wellington, New Zealand, pp. 59-70.
Barnett, J. and S. O’Neill, 2010: Maladaptation. Global Environmental Change, 20(2),
211-213.
Barnett, J., E. Waters, S. Pendergast, and A. Puleston, 2013: Barriers to Adaptation to
Sea-Level Rise. The University of Melbourne and National Climate Change
Adaptation Research Facility (NCCARF), NCCARF, Griffith University, Gold Coast
Campus, QLD, Australia, 85 pp.
Barton, A.B. and J.R. Argue, 2009: Integrated urban water management for residential
areas: a reuse model. Water Science and Technology, 60(3), 813-823.
Basher, L., S. Elliot, A. Hughes, A. Tait, M. Page, B. Rosser, I. McIvor, G. Douglas, and
H. Jones, 2012: Impacts of Climate Change on Erosion and Erosion Control
Methods – A Critical Review. MPI Technical Paper No. 2012/45, Ministry for
Primary Industries, Wellington, New Zealand, 208 pp.
Bates, B.C., P. Hope, B. Ryan, I. Smith, and S. Charles, 2008: Key findings from the
Indian Ocean Climate Initiative and their impact on policy development in
Australia. Climatic Change, 89(3-4), 339-354.
Bates, B.C., R.E. Chandler, S.P. Charles, and E.P. Campbell, 2010: Assessment of
apparent nonstationarity in time series of annual inflow, daily precipitation, and
atmospheric circulation indices: a case study from southwest Western Australia.
Water Resources Research, 46(3), W00H02, doi:10.1029/2010WR009509.
Bates, B.C., S. Bunn, P. Baker, M. Cox, A. Hopkins, B. Humphreys, S. Lakes, G. Willgoose,
and B. Young, 2011: National Climate Change Adaptation Research Plan for
Freshwater Biodiversity. National Climate Change Adaptation Research Facility
(NCCARF), Griffith University, Gold Coast Campus, QLD, Australia 68 pp.
Battaglene, S.C., C. Carter, A.J. Hobday, V. Lyne, and B. Nowak, 2008: Scoping Study
into Adaptation of the Tasmanian Salmonid Aquaculture Industry to Potential
Impacts of Climate Change. National Agriculture & Climate Change Action Plan:
Implementation Programme Report, Tasmanian Aquaculture and Fisheries
Institute (TAFI), University of Tasmania, and CSIRO Marine and Atmospheric
Research, TAFI, Hobart, TAS, Australia, 84 pp.
Battaglia, M., J. Bruce, C. Brack, and T. Baker, 2009: Climate Change and Australia’s
Plantation Estate: Analysis of Vulnerability and Preliminary Investigation of
Adaptation Options. Forest & Wood Products Australia Ltd., Melbourne, VIC,
Australia, 125 pp.
Baum, S., S. Horton, D.L. Choy, and B. Gleeson, 2009: Climate Change, Health Impacts
and Urban Adaptability: Case Study of Gold Coast City. Research Monograph
11, Griffith University, Urban Research Program, Brisbane, QLD, Australia,
67 pp.
Baynes, T., A. Herr, A. Langston, and H. Schandl, 2012: Coastal Climate Risk Project
Milestone 1 Final Report to the Australian Department of Climate Change and
Energy Efficiency. Prepared for the Department of Climate Change and Energy
Efficiency (DCCEE) by the Commonwealth Scientific and Industrial Research
Organization (CSIRO), DCCEE, Canberra, ACT, Australia, 102 pp.
Becken, S., 2011: A critical review of tourism and oil. Annals of Tourism Research,
38(2), 359-379.
Becken, S. and R. Clapcott, 2011: National tourism policy for climate change. Journal
of Policy Research in Tourism, Leisure and Events, 3(1), 1-17.
Bedford, R. and C. Bedford, 2010: International migration and climate change: a
post-Copenhagen perspective on options for Kiribati and Tuvalu. In: Climate
Change and Migration: South Pacific Perspectives [Burson, B. (ed.)]. Institute
of Policy Studies, Victoria University of Wellington, Wellington, New Zealand,
pp. 89-134.
Bedford, R. and J. Campbell, 2013: Migration and climate change in Oceania. In: People
on the Move in a Changing Climate: The Regional Impact of Environmental
Change on Migration [Piguet, E. and F. Laczko (eds.)]. Global Migration Issues,
Vol. 2, International Organization for Migration (IOM), Geneva, Switzerland,
and Springer, Dordrecht, Netherlands, pp. 177-204.
Beebe, N.W., R.D. Cooper, P. Mottram, and A.W. Sweenet, 2009: Australia’s dengue
risk driven by human adaptation to climate change. PLoS Neglected Tropical
Diseases, 3(5), e429, doi:10.1371/journal.pntd.0000429.
Beentjes, M.P. and J.A. Renwick, 2001: The relationship between red cod, Pseudophycis
bachus, recruitment and environmental variables in New Zealand. Environmental
Biology of Fishes, 61(3), 315-328.
Beggs, P.J. and C.M. Bennett, 2011: Climate change, aeroallergens, natural particulates
and human health in Australia: state of science and policy. Asia Pacific Journal
of Public Health, 23(2 Suppl.), 46S-53S.
Béguin, A., S. Hales, J. Rocklöv, C. Åström, V.R. Louis, and R. Sauerborn, 2011: The
opposing effects of climate change and socio-economic development on the
global distribution of malaria. Global Environmental Change, 21(4), 1209-
1214.
Bell, M., G. Blick, O. Parkyn, P. Rodway, and P. Vowles, 2010: Challenges and Choices:
Modelling New Zealand’s Long-term Fiscal Position. Working Paper 10/01, New
Zealand Treasury, Wellington, New Zealand, 85 pp.
Bennett, C.M., K.B.G. Dear, and A.J. McMichael, 2013: Shifts in the seasonal
distribution of deaths in Australia, 1968-2007. International Journal of
Biometeorology, doi: 10.1007/s00484-013-0663-x.
1416
Chapter 25 Australasia
25
Bergin, A. and J. Townsend, 2007: A Change in Climate for the Australian Defence Force.
Special Report 2007-7, Australian Strategic Policy Institute (ASPI), Canberra,
ACT, Australia, 11 pp.
Berry, H.L., J.R.A. Butler, C.P. Burgess, U.G. King, K. Tsey, Y.L. Cadet-James, C.W. Rigby,
and B. Raphael, 2010: Mind, body, spirit: co-benefits for mental health from
climate change adaptation and caring for country in remote Aboriginal
Australian communities. New South Wales Public Health Bulletin, 21(5-6), 139-
145.
Berry, S. and J. Vella, 2010: Planning Controls and Property Rights Striking the
Balance. Resource Management Law Association Roadshow 2010, Simon Berry
Environmental Law, Auckland, New Zealand, 64 pp.
Betts, R.A., O. Boucher, M. Collins, P.M. Cox, P.D. Falloon, N. Gedney, D.L. Hemming,
C. Huntingford, C.D. Jones, D.M.H. Sexton, and M.J. Webb, 2007: Projected
increase in continental runoff due to plant responses to increasing carbon dioxide.
Nature, 448, 1037-1041.
Bevin, S., 2007: Economic Impact of the 2007 East Coast Drought on the Sheep and
Beef Sector. Report prepared by Economic Solutions Ltd., Napier for the Ministry
of Agriculture and Forestry, Wellington, New Zealand, 14 pp.
Bi, P. and K.A. Parton, 2008: Effect of climate change on Australian rural and remote
regions: what do we know and what do we need to know? Australian Journal
of Rural Health, 16(1), 2-4.
Bicknell, S. and P. McManus, 2006: The canary in the coalmine: Australian ski resorts
and their response to climate change. Geographical Research, 44(4), 386-400.
Bird, D.W., R.B. Bird, and C.H. Parker, 2005: Aboriginal burning regimes and hunting
strategies in Australia’s western desert. Human Ecology, 33(4), 443-464.
Blanchette, C.A., E.A. Wieters, B.R. Broitman, B.P. Kinlan, and D.R. Schiel, 2009:
Trophic structure and diversity in rocky intertidal upwelling ecosystems: a
comparison of community patterns across California, Chile, South Africa and
New Zealand. Progress in Oceanography, 83(1-4), 107-116.
BMT WBM, 2011: Kakadu: Vulnerability to Climate Change Impacts. Report to the
Australian Government Department of Climate Change and Energy Efficiency
(DCCEE), prepared by BMT WBM Pty Ltd.-Brisbane with expert input from
Melaleuca Enterprises, Coastal Zone Management Pty Ltd., and Jacqueline
Robinson (UQ), DCCEE, Canberra, ACT, Australia, 226 pp.
Bohensky, E., J.R.A. Butler, R. Costanza, I. Bohnet, A. Delisle, K. Fabricius, M. Gooch,
I. Kubiszewski, G. Lukacs, P. Pert, and E. Wolanski, 2011: Future makers or future
takers? A scenario analysis of climate change and the Great Barrier Reef. Global
Environmental Change, 21(3), 876-893.
BoM, 2009: The Exceptional January-February 2009 Heatwave in South-Eastern
Australia. Special Climate Statement 17, National Climate Centre, Bureau of
Meteorology (BoM), Melbourne, VIC, Australia,11 pp.
BoM, 2013: Special Climate Statement 43 – Extreme Heat in January 2013. Bureau
of Meteorology (BoM), Melbourne, VIC, Australia, 23 pp.
Bond, N., P. Lake, and A. Arthington, 2008: The impacts of drought on freshwater
ecosystems: an Australian perspective. Hydrobiologia, 600(1), 3-16.
Bond, N., J. Thomson, P. Reich, and J. Stein, 2011: Using species distribution models
to infer potential climate change-induced range shifts of freshwater fish in
south-eastern Australia. Marine and Freshwater Research, 62(9), 1043-1061.
Booth, K. and S. Williams, 2012: Is insurance an under-utilised mechanism in climate
change adaptation? The case of bushfire management in Tasmania. Australian
Journal of Emergency Management, 27(4), 38-45.
Booth, T.H., M.U.F. Kirschbaum, and M. Battaglia, 2010: Forestry. In: Adapting
Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and
Fisheries for the Future [Stokes, C. and S. Howden (eds.)]. Commonwealth
Scientific and Industrial Research Organization (CSIRO) Publishing, Collingwood,
VIC, Australia, pp. 137-152.
Botterill, L.C. and S. Dovers, 2013: Drought and water in the Murray-Darling Basin:
from disaster policy to adaptation. In: Natural Disasters and Adaptation to
Climate Change [Boulter, S., J. Palutikof, D. Karoly, and D. Guitart (eds.)].
Cambridge University Press, Cambridge, UK, pp. 100-114.
Bourdôt, G.W., S.L. Lamoureaux, M.S. Watt, L.K. Manning, and D.J. Kriticos, 2012:
The potential global distribution of the invasive weed Nassella neesiana under
current and future climates. Biological Invasions, 14(8), 1545-1556.
Bowman, D., B.P. Murphy, and D.S. Banfai, 2010: Has global environmental change
caused monsoon rainforests to expand in the Australian monsoon tropics?
Landscape Ecology, 25(8), 1247-1260.
Bradshaw, C.J.A., X. Giam, and N.S. Sodhi, 2010: Evaluating the relative environmental
impact of countries. PLoS ONE, 5(5), e10440, doi: 10.1371/journal.pone.
0010440.
Bradshaw, C.J.A., D.M.J.S. Bowman, N.R. Bond, B.P. Murphy, A.D. Moore, D.A. Fordham,
R. Thackway, M.J. Lawes, H. McCallum, S.D. Gregory, R.C. Dalal, M.M. Boer, A.J.J.
Lynch, R.A. Bradstock, B.W. Brook, B.K. Henry, L.P. Hunt, D.O. Fisher, D. Hunter,
C.N. Johnson, D.A. Keith, E.C. Lefroy, T.D. Penman, W. Meyer, J.R. Thomson, C.M.
Thornton, J. Van Der Wal, D. Williams, L. Keniger, and A. Specht, 2013: Brave
new green world – consequences of a carbon economy for the conservation of
Australian biodiversity. Biological Conservation, 161, 71-90.
Bradstock, R.A., 2010: A biogeographic model of fire regimes in Australia: current
and future implications. Global Ecology and Biogeography, 19(2), 145-158.
BRANZ, 2007: An Assessment of the Need to Adapt Buildings for the Unavoidable
Consequences of Climate Change. Report by the Building Research Association
of New Zealand (BRANZ) Ltd., for the Australian Greenhouse Office,
Department of the Environment and Water Resources, Canberra, ACT, Australia,
81 pp.
Bright, J., H. Rutter, J. Dommisse, R. Woods, A. Tait, B. Mullan, J. Hendrix, and J.
Diettrich, 2008: Projected Effects of Climate Change on Water Supply Reliability
in Mid-Canterbury. Report C08120/1 by Aqualine Research Ltd., for the Ministry
of Agriculture and Forestry (MAF), Wellington, New Zealand, 45 pp.
Britton, E., S. Hales, K. Venugopal, and M.G. Baker, 2010a: Positive association
between ambient temperature and salmonellosis notifications in New Zealand,
1965-2006. Australian and New Zealand Journal of Public Health, 34(2), 126-
129.
Britton, E., S. Hales, K. Venugopal, and M.G. Baker, 2010b: The impact of climate
variability and change on cryptosporidiosis and giardiasis rates in New Zealand.
Journal of Water and Health, 8(3), 561-571.
Britton, R., 2010: Coastal Adaptation to Climate Change. Report on Local Government
Planning Practice and Limitations to Adaptation. National Institute of Water
and Atmospheric Research (NIWA), Auckland, New Zealand, 51 pp.
Britton, R., J. Dahm, H. Rouse, T. Hume, R. Bell, and P. Blackett, 2011: Coastal Adaptation
to Climate Change: Pathways to Change. Report prepared as part of the Coastal
Adaptation to Climate Change Project, National Institute of Water and
Atmospheric Research (NIWA), Auckland, New Zealand, 106 pp.
Brockerhoff, E.G., H. Jactel, J.A. Parrotta, C.P. Quine, and J. Sayer, 2008: Plantation
forests and biodiversity: oxymoron or opportunity? Biodiversity Conservation,
17, 925-951.
Brown, A.E., L. Zhang, T.A. McMahon, A.W. Western, and R.A. Vertessy, 2005: A review
of paired catchment studies for determining changes in water yield resulting
from alterations in vegetation. Journal of Hydrology, 310, 28-61.
Brown, R.R. and M. Farrelly, 2009: Challenges ahead: social and institutional factors
influencing sustainable urban stormwater management in Australia. Water
Science and Technology, 59(4), 653-660.
Brown, R.R., M. Farrelly, and N. Keath, 2009: Practitioner perceptions of social and
institutional barriers to advancing a diverse water source approach in Australia.
International Journal of Water Resources Development, 25(1), 15-28.
Brunckhorst, D., I. Reeve, P. Morley, M. Coleman, E. Barclay, J. McNeill, R. Stayner, R.
Glencross-Grant, J. Thompson, and L. Thompson, 2011: Hunter & Central Coasts
New South Wales – Vulnerability to Climate Change Impacts. A report to the
Australian Government Department of Climate Change and Energy Efficiency
(DCCEE) by the Institute for Rural Futures, University of New England, DCCEE,
Canberra, ACT, Australia, 158 pp.
Bruno, J.F., E.R. Selig, K.S. Casey, C.A. Page, B.L. Willis, C.D. Harvell, H. Sweatman,
and A.M. Melendy, 2007: Thermal stress and coral cover as drivers of coral
disease outbreaks. PLoS Biology, 5(6), 1220-1227.
Bryan, B.A., D. King, and E.L. Wang, 2010: Biofuels agriculture: landscape-scale trade-
offs between fuel, economics, carbon, energy, food, and fiber. Global Change
Biology Bioenergy, 2(6), 330-345.
BSC, 2010: Draft Coastal Zone Management Plan for Byron Shire Coastline. Part A:
The Plan. Byron Shire Council, Mullumbimby, NSW, Australia, 413 pp.
Buckley, R., 2008: Misperceptions of climate change damage coastal tourism: case
study of Byron Bay, Australia. Tourism Review International, 12, 71-88.
Burbidge, A.A., M. Byrne, D. Coates, S.T. Garnett, S. Harris, M.W. Hayward, T.G. Martin,
E. McDonald-Madden, N.J. Mitchell, S. Nally, and S.A. Setterfield, 2011: Is
Australia ready for assisted colonization? Policy changes required to facilitate
translocations under climate change. Pacific Conservation Biology, 17(3), 259-
269.
Burgess, C.P., F.H. Johnston, H. Berry, J. McDonnell, D. Yibarbuk, C. Gunabarra, A.
Mileran, and R. Bailie, 2009: Healthy country healthy people: superior indigenous
health outcomes are associated with ‘caring for country’. Medical Journal of
Australia, 190(10), 567-572.
1417
Australasia Chapter 25
25
Burgette, R.J., C.S. Watson, J.A. Church, N.J. White, P. Tregoning, and R. Coleman,
2013: Characterizing and minimizing the effects of noise in tide gauge time
series: relative and geocentric sea level rise around Australia. Geophysical
Journal International, 194(2), 719-736.
Burroughs, S., 2010: Sustainable housing for Indigenous populations in remote
regions: a case study from Nguiu, Northern Australia. In: CESB 10: Central
Europe towards Sustainable Building: From Theory to Practice. Keynote Lectures.
Czech Technical University, Prague, Czech Republic, pp. 11-18.
Burton, P. and J. Mustelin, 2013: Planning for climate change: is greater public
participation the key to success? Urban Policy and Research, 31(4), 399-415.
Butcher, 2009: Regional and National Impacts of the 2007-2009 Drought. Report
by Butcher Partners Ltd., for the Ministry of Agriculture and Forestry (MAF)
Policy, Wellington, New Zealand, 32 pp.
Buultjens, J., D. Gale, and N.E. White, 2010: Synergies between Australian Indigenous
tourism and ecotourism: possibilities and problems for future development.
Journal of Sustainable Tourism, 18(4), 497-513.
Byron, N., 2011: What can the Murray-Darling Basin Plan achieve? Will it be enough?
In: Basin Futures: Water Reform in the Murray-Darling Basin [Grafton, R.Q. and
D. Connell (eds.)]. The Australian National University (ANU), ANU E-Press,
Canberra, ACT, Australia, pp. 385-398.
Cai, W. and T. Cowan, 2006: SAM and regional rainfall in IPCC AR4 models: can
anthropogenic forcing account for southwest Western Australian winter rainfall
reduction? Geophysical Research Letters, 33(24), L24708, doi:10.1029/
2006GL028037.
Cai, W. and T. Cowan, 2008: Evidence of impacts from rising temperature on inflows
to the Murray-Darling Basin. Geophysical Research Letters, 35(7), L07701,
doi:10.1029/2008GL033390.
Cai, W., T. Cowan, and M. Raupach, 2009a: Positive Indian Ocean Dipole events
precondition southeast Australia bushfires. Geophysical Research Letters,
36(19), L19710, doi:10.1029/2009GL039902.
Cai, W., A. Sullivan, and T. Cowan, 2009b: Climate change contributes to more frequent
consecutive positive Indian Ocean Dipole events. Geophysical Research Letters,
36(23), L23704, doi:10.1029/2009GL040163.
Cai, W., P. van Rensch, and T. Cowan, 2011: Influence of global-scale variability on the
subtropical ridge over Southeast Australia. Journal of Climate, 24(23), 6035-6053.
Cai, W., P. van Rensch, and T. Cowan, 2012: The 2011 southeast Queensland floods:
a confirmation of a negative Pacific Decadal Oscillation phase? Geophysical
Research Letters, 39(8), L08702, doi:10.1029/2011GL050820.
Calder, J.A. and J.B. Kirkpatrick, 2008: Climate change and other factors influencing
the decline of the Tasmanian cider gum (Eucalyptus gunnii). Australian Journal
of Botany, 56(8), 684-692.
Callaghan, J. and S.B. Power, 2011: Variability and decline in the number of severe
tropical cyclones making landfall over eastern Australia since the late nineteenth
century. Climate Dynamics, 37(3-4), 647-662.
Cameron, P.A., B. Mitra, M. Fitzgerald, C.D. Scheinkestel, A. Stripp, C. Batey, L.
Niggemeyer, M. Truesdale, P. Holman, R. Mehra, J. Wasiak, and H. Cleland, 2009:
Black Saturday: the immediate impact of the February 2009 bushfires in Victoria,
Australia. The Medical Journal of Australia, 191, 11-16.
Campbell, A., 2008: Managing Australian Landscapes in a Changing Climate: A
Climate Change Primer for Regional Natural Resource Management Bodies.
Department of Climate Change and Energy Efficiency, Canberra, ACT, Australia,
46 pp.
Campbell, D., M. Stafford Smith, J. Davies, P. Kuipers, J. Wakerman, and M.J. McGregor,
2008: Responding to health impacts of climate change in the Australian desert.
Rural and Remote Health, 8(3), 1008, www.rrh.org.au/publishedarticles/
article_print_1008.pdf.
Carey-Smith, T., S. Dean, J. Vial, and C. Thompson, 2010: Changes in precipitation
extremes for New Zealand: climate model predictions. Weather and Climate,
30, 23-48.
Carr, M.K.V. and J.W. Knox, 2011: The water relations and irrigation requirements of
sugarcane (Saccharum officinarum): a review. Experimental Agriculture, 47(1),
1-25.
Carswell, F., G. Harmsworth, R. Kirikiri, I. Turney, and S. Kerr, 2002: A Framework for
Engagement of ori Landowners in “Carbon Farming” Using Indigenous
Forest Regeneration. Landcare Research Contract Report: LC0102/116,
Prepared for Non-Specific-Output Funding (NSOF) by Landcare Research New
Zealand Ltd., Lincoln, New Zealand, 24 pp.
Carter, T.R., 2013: Agricultural impacts: multi-model yield projections. Nature Climate
Change, 3(9), 784-786.
Cary, G., D. Lindenmayer, and S. Dovers (eds.), 2003: Australia Burning Fire Ecology,
Policy and Management Issues. Commonwealth Scientific and Industrial
Research Organization (CSIRO) Publishing, Collingwood, VIC, Australia, 268 pp.
CCC, 2010: Christchurch City Council Climate Smart Strategy 2010-2025. Christchurch
City Council (CCC), Christchurch, New Zealand, 46 pp.
Chakraborty, S., J. Luck, G. Hollaway, G. Fitzgerald, and N. White, 2011: Rust-proofing
wheat for a changing climate. Euphytica, 179, 19-32.
Chambers, L.E. and G.M. Griffiths, 2008: The changing nature of temperature
extremes in Australia and New Zealand. Australian Meteorological Magazine,
57, 13-55
Chambers, L.E., C.A. Devney, B.C. Congdon, N. Dunlop, E.J. Woehler, and P. Dann,
2011: Observed and predicted effects of climate on Australian seabirds. Emu,
111(3), 235-251.
Chambers, L.E., R. Altwegg, C. Barbraud, P. Barnard, L.J. Beaumont, R.J.M. Crawford,
J.M. Durant, L. Hughes, M.R. Keatley, M. Low, L.P.C. Morellato, E.S. Poloczanska,
V. Ruoppolo, R.E.T. Vanstreels, E.J. Woehler, and A.C. Wolfaardt, 2013a:
Phenological changes in the Southern Hemisphere. PLoS ONE, 8(10), e75514,
doi: 10.1371/journal.pone.0075514.
Chambers, L.E., L. Beaumont, and I. Hudson, 2013b: Continental scale analysis of
bird migration timing: influences of climate and life history traits; a generalised
mixture model clustering approach. International Journal of Biometeorology,
doi: 10.1007/s00484-013-0707-2.
Chessman, B.C., 2009: Climatic changes and 13-year trends in stream macroinvertebrate
assemblages in New South Wales, Australia. Global Change Biology, 15(11),
2791-2802.
Cheung, W.W.L., J.J. Meeuwig, M. Feng, E. Harvey, V.W.Y. Lam, T. Langlois, D. Slawinski,
C.J. Sun, and D. Pauly, 2012: Climate-change induced tropicalisation of marine
communities in Western Australia. Marine and Freshwater Research, 63(5),
415-427.
Chiew, F., 2006: Estimation of rainfall elasticity of streamflow in Australia.
Hydrological Sciences Journal/Journal Des Sciences Hydrologiques, 51(4), 613-
625.
Chiew, F. and T.A. McMahon, 2002: Global ENSO-streamflow teleconnection, streamflow
forecasting and interannual variability. Hydrological Sciences Journal/Journal
Des Sciences Hydrologiques, 47(3), 505-522.
Chiew, F. and I.P. Prosser, 2011: Water and climate. In: Water: Science and Solutions
for Australia [Prosser, I. (ed.)]. Commonwealth Scientific and Industrial Research
Organization (CSIRO) Publishing, Collingwood, VIC, Australia, pp. 29-46.
Chiew, F., J. Teng, J. Vaze, D.A. Post, J.M. Perraud, D.G.C. Kirono, and N.R. Viney, 2009:
Estimating climate change impact on runoff across southeast Australia: method,
results, and implications of the modeling method. Water Resources Research,
45(10), W10414, doi:10.1029/2008WR007338.
Chiew, F., W.J. Young, W. Cai, and J. Teng, 2011: Current drought and future
hydroclimate projections in southeast Australia and implications for water
resources management. Stochastic Environmental Research and Risk Assessment,
25(4), 601-612.
Chiew, F., N.J. Potter, J. Vaze, C. Petheram, L. Zhang, J. Teng, and D.A. David, 2013:
Observed hydrologic non-stationarity in far south-eastern Australia: implications
and future modelling predictions. Stochastic Environmental Research and Risk
Assessment, 28(1), 3-15.
Chinn, T.J., 2001: Distribution of the glacial water resources of New Zealand. Journal
of Hydrology (NZ), 40(2), 139-187.
Chinn, T.J., S. Winkler, M.J. Salinger, and N. Haakensen, 2005: Recent glacier advances
in Norway and New Zealand: a comparison of their glaciological and
meteorological causes. Geografiska Annaler: Series A, Physical Geography,
87(1), 141-157.
Chinn, T.J., B.B. Fitzharris, A. Willsman, and M.J. Salinger, 2012: Annual ice volume
changes 1976-2008 for the New Zealand Southern Alps. Global and Planetary
Change, 92-93, 105-118.
Chiswell, S.M. and J.D. Booth, 2005: Distribution of mid- and late-stage Jasus
edwardsii phyllosomas: implications for larval recruitment processes. New
Zealand Journal of Marine and Freshwater Research, 39(5), 1157-1170.
City of Adelaide, 2012: The City of Adelaide Strategic Plan 2012-2016. Adelaide
City Council, Adelaide, SA, Australia, 28 pp.
City of Melbourne, 2009: City of Melbourne Climate Change Adaptation Strategy.
Prepared by Maunsell Australia Pty Ltd., City of Melbourne, Melbourne, VIC,
Australia, 128 pp.
City of Port Phillip, 2010: Greening Port Phillip, an Urban Forest Approach. Port
Phillip City Council, St. Kilda, VIC, Australia, 54 pp.
1418
Chapter 25 Australasia
25
Clark, A. and A. Tait, 2008: Drought, Agricultural Production & Climate Change A Way
Forward to a Better Understanding. NIWA Client Report WLG2008-33, Prepared
for the Ministry for Agriculture and Fisheries by the National Institute of Water
& Atmospheric Research (NIWA), NIWA, Wellington, New Zealand, 71 pp.
Clark, A., B. Mullan, and A. Porteous, 2011: Scenarios of Regional Drought under
Climate Change. NIWA Client Report WLG2012-32, Prepared for Ministry for
Agriculture and Fisheries by the National Institute of Water and Atmospheric
Research (NIWA), NIWA, Wellington, New Zealand, 135 pp.
Clark, A., R. Nottage, and D. Hansford (eds.), 2012: Impacts of Climate Change on
Land-based Sectors and Adaptation Options. Stakeholder Report. Ministry of
Primary Industries, Wellington, New Zealand, 74 pp.
Clarke, H.G., P.L. Smith, and A.J. Pitman, 2011: Regional signatures of future fire
weather over eastern Australia from global climate models. International Journal
of Wildland Fire, 20(4), 550-562.
Clarke, H.G., C. Lucas, and P.L. Smith, 2012: Changes in Australian fire weather between
1973 and 2010. International Journal of Climatology, 33(4), 931-944.
Clayton, H., S. Dovers, and P. Harris, 2011: Natural Resource Management Policy and
Planning in Australia. Synthesis of Broad Issues and Opportunities: Document
I, Literature Review: Document II, and Workshop Report: Document III. HC
Coombs Policy Forum-Fenner School of Environment and Society NRM initiative,
The Australian National University (ANU), Canberra, ACT, Australia, 37 pp.
Clothier, B., A. Hall, and S. Green, 2012: Horticulture: adapting the horticultural and
vegetable industries to climate change. In: Enhanced Climate Change Impact
and Adaptation Evaluation: A Comprehensive Analysis of New Zealand’s Land-
Based Primary Sectors [Clark, A. and Nottage, R. (eds.)]. Ministry for Primary
Industries, Wellington, New Zealand, pp. 237-292.
Clucas, R., 2011: Long-term population trends of Sooty Shearwater (Puffinus griseus)
revealed by hunt success. Ecological Applications, 21(4), 1308-1326.
COAG, 2007: National Climate Change Adaptation Framework. Council of Australian
Governments (COAG), Department of the Prime Minister and Cabinet,
Canberra, ACT, Australia, 27 pp.
Cobon, D.H., G.S. Stone, J.O. Carter, J.C. Scanlan, N.R. Toombs, X. Zhang, J. Willcocks,
and G.M. McKeon, 2009: The climate change risk management matrix for the
grazing industry of northern Australia. The Rangeland Journal, 31(1), 31-49.
Cock, M.J.W., J.C. van Lenteren, J. Brodeur, B.I.P. Barratt, F. Bigler, K. Bolckmans, F.L.
Cônzoli, F. Haas, P.G. Mason, and J.G.P. Parra, 2010: Do new access and benefit
sharing procedures under the convention on biological diversity threaten the
future of biological control? BioControl, 55, 199-218.
Cocklin, C. and J. Dibden, 2009: Systems in peril: climate change, agriculture and
biodiversity in Australia. IOP Conference Series: Earth and Environmental
Science, 8(1), 012013, doi:10.1088/1755-1315/8/1/012013.
Collins, D. and R. Kearns, 2010: “It’s a gestalt experience”: landscape values and
development pressure in Hawke’s Bay, New Zealand. Geoforum, 41(3), 435-
446.
Connell, D., 2007: Water Politics in the Murray-Darling Basin. The Federation Press,
Sydney, NSW, Australia, 241 pp.
Connell, D. and Q. Grafton (eds.), 2011: Basin Futures: Water Reform in the Murray-
Darling Basin. The Australian National University (ANU) E-Press, Canberra, ACT,
Australia, 477 pp.
Connor, J., K. Schwabe, D. King, D. Kaczan, and M. Kirby, 2009: Impacts of climate
change on lower Murray irrigation. Australian Journal of Agricultural and
Resources Economics, 53(3), 437-456.
Cooper, T.F., G. De’ath, K.E. Fabricius, and J.M. Lough, 2008: Declining coral calcification
in massive Porites in two nearshore regions of the northern Great Barrier Reef.
Global Change Biology, 14(3), 529-538.
Cooper, T.F., R.A. O’Leary, and J.M. Lough, 2012: Growth of Western Australian corals
in the Anthropocene. Science, 335(6068), 593-596.
Corkhill, J.R., 2013: Claimed property right does not hold water. Australian Law
Journal, 87, 49-57.
Cornish, P.M. and R.A. Vertessy, 2001: Forest age-induced changes in evapotranspiration
and water yield in a eucalypt forest. Journal of Hydrology, 242(1-2), 43-63.
Cottrell, B., C. Insley, R. Meade, and J. West, 2004: Report of the Climate Change
Māori Issues Group. New Zealand Climate Change Office, Wellington, New
Zealand, 27 pp.
Coutts, A., J. Beringer, and N. Tapper, 2010: Changing urban climate and CO
2
emissions:
implications for the development of policies for sustainable cities. Urban Policy
and Research, 28(1), 27-47.
Crase, L., 2011: The fallout to the Guide to the Proposed Basin Plan. Australian
Journal of Public Administration, 70(1), 84-93.
Crompton, R.P. and K.J. McAneney, 2008: Normalised Australian insured losses from
meteorological hazards: 1967-2006. Environmental Science & Policy, 11(5),
371-378.
Crompton, R.P., K.J. McAneney, K. Chen, R.A. Pielke Jr., and K. Haynes, 2010: Influence
of location, population, and climate on building damage and fatalities due to
Australian Bushfire: 1925-2009. Weather, Climate, and Society, 2(4), 300-310.
Crompton, R.P., K.J. McAneney, K. Chen, R.A. Pielke Jr., and K. Haynes, 2011:
Response to Nicholls (2011) on “Influence of location, population and climate
on building damage and fatalities due to Australian Bushfire: 1925-2009”.
Weather, Climate, and Society, 3(1), 63-66, doi:10.1175/WCAS-D-11-00002.
Crook, J.A., L.A. Jones, P.M. Forster, and R. Crook, 2011: Climate change impacts on
future photovoltaic and concentrated solar power energy output. Energy &
Environmental Science, 4(9), 3101-3109.
Crosbie, R.S., J.L. McCallum, G.R. Walker, and F. Chiew, 2010: Modelling climate-
change impacts on groundwater recharge in the Murray-Darling Basin,
Australia. Hydrogeology Journal, 18(7), 1639-1656.
Crosbie, R.S., T. Pickett, F.S. Mpelasoka, G. Hodfgson, S.P. Charles, and O. Barron,
2012: An assessment of the climate change impacts on groundwater recharge
at a continental scale using a probabilistic approach with an ensemble of
GCMs. Climatic Change, 117(1-2), 41-53.
Crossman, N.D., B.A. Bryan, and D.M. Summers, 2012: Identifying priority areas for
reducing species vulnerability to climate change. Diversity and Distributions,
18(1), 60-72.
Cruz, F., A. Pitman, and Y. Wang, 2010: Can the stomatal response to higher
atmospheric carbon dioxide explain the unusual temperatures during the 2002
Murray-Darling Basin drought? Journal of Geophysical Research: Atmospheres,
115(D2), D02101, doi:10.1029/2009JD012767.
CSIRO, 2008: Water Availability in the Murray-Darling Basin. A Report to the
Australian Government from the CSIRO Murray-Darling Basin Sustainable Yields
Project, Commonwealth Scientific and Industrial Research Organization
(CSIRO), CSIRO Publishing, Collingwood, VIC, Australia, 68 pp.
CSIRO, 2009: Surface Water Yields in South-West Western Australia. A Report to the
Australian Government from the CSIRO South-West Western Australia Sustainable
Yields Project, Commonwealth Scientific and Industrial Research Organization
(CSIRO) Water for a Healthy Country Flagship, CSIRO Publishing, Collingwood,
VIC, Australia, 195 pp.
CSIRO, 2010: Climate Variability and Change in South-Eastern Australia: A Synthesis
of Findings from Phase 1 of the South Eastern Australian Climate Initiative
(SEACI). Prepared for the South Eastern Australian Climate Initiative Steering
Committee by the Commonwealth Scientific and Industrial Research Organization
(CSIRO) Water for a Healthy Country Flagship, CSIRO Publishing, Collingwood,
VIC, Australia, 31 pp.
CSIRO, 2011: Barriers to Effective Climate Change Adaptation: Submissions. CSIRO
Submission to Productivity Commission Issues Paper, No. 40 (CSIRO Submission
11/432), Commonwealth Scientific and Industrial Research Organization
(CSIRO) Climate Adaptation Flagship, Productivity Commission, Melbourne, VIC,
Australia, 30 pp., www.pc.gov.au/projects/inquiry/climate-change-adaptation/
submissions.
CSIRO, 2012: Climate and Water Availability in South-Eastern Australia: A Synthesis
of Findings from Phase 2 of the South Eastern Australian Climate Initiative
(SEACI). Commonwealth Scientific and Industrial Research Organization
(CSIRO) Water for a Healthy Country Flagship, CSIRO Publishing, Collingwood,
VIC, Australia, 41 pp.
CSIRO and BoM, 2007: Climate Change in Australia: Technical Report, 2007.
Commonwealth Scientific and Industrial Research Organization (CSIRO) and
Bureau of Meteorology (BoM), CSIRO Publishing, Collingwood, VIC, Australia,
148 pp.
CSIRO and BoM, 2012: State of the Climate 2012. Commonwealth Scientific and
Industrial Research Organization (CSIRO), CSIRO Publishing, Collingwood, VIC,
Australia, 12 pp., www.csiro.au/Outcomes/Climate/Understanding/State-of-the-
Climate-2012.aspx.
CSIRO, Maunsell Australia Pty Ltd., and Phillips Fox, 2007: Infrastructure and Climate
Change Risk Assessment for Victoria. Report prepared for Victorian Government
by the Commonwealth Scientific and Industrial Research Organization (CSIRO)
Marine and Atmospheric Research, Aspendale, VIC, Australia, 72 pp.
Cullen, B.R. and R.J. Eckard, 2011: Impacts of future climate scenarios on the balance
between productivity and total greenhouse gas emissions from pasture based
dairy systems in south-eastern Australia. Animal Feed Science and Technology,
166-167, 721-735.
1419
Australasia Chapter 25
25
Cullen, J.M., L.E. Chambers, P.C. Coutin, and P. Dann, 2009: Predicting onset and success
of breeding in Little Penguins (Eudyptula minor) from ocean temperatures.
Marine Ecology Progress Series, 378, 269-278.
DAFF, 2012: Australias Agriculture, Fisheries and Forestry at a Glance 2012. Department
of Agriculture, Fisheries and Forestry (DAFF), Canberra, ACT, Australia, 131 pp.
Dalton, S.J., S. Godwin, S.D.A. Smith, and L. Pereg, 2010: Australian subtropical white
syndrome: a transmissible, temperature-dependent coral disease. Marine and
Freshwater Research, 61(3), 342-350.
Darbyshire, R., L. Webb, I. Goodwin, and E.W.R. Barlow, 2013: Evaluation of recent
trends in Australian pome fruit spring phenology. International Journal of
Biometeorology, 57(3), 409-421.
Davis, J., A. Pavlova, R. Thompson, and P. Sunnucks, 2013: Evolutionary refugia and
ecological refuges: key concepts for conserving Australian arid zone freshwater
biodiversity under climate change. Global Change Biology, 19, 1970-1984.
Davies, J., D. Campbell, M. Campbell, J. Douglas, H. Hueneke, M. LaFlamme, D. Pearson,
K. Preuss, J. Walker, and F.J. Walsh, 2011: Attention to four key principles can
promote health outcomes from desert Aboriginal land management. The
Rangeland Journal, 33(4), 417-431.
Davies, P.M., 2010: Climate change implications for river restoration in global
biodiversity hotspots. Restoration Ecology, 18(3), 261-268.
DCC, 2009: Climate Change Risks to Australias Coast: A First-Pass National Assessment.
Department of Climate Change (DCC), Government of Australia, Canberra, ACT,
Australia, 168 pp.
DCCEE, 2010: Developing a National Coastal Adaptation Agenda. A Report on the
National Climate Change Forum: Adaptation Priorities for Australia’s Coasts,
Adelaide, 18-19 February 2010. Department of Climate Change and Energy
Efficiency (DCCEE), Canberra, ACT, Australia, 48 pp.
DCCEE, 2011: Climate Change Risks to Coastal Buildings and Infrastructure – A
Supplement to the First Pass National Assessment. Department of Climate
Change and Energy Efficiency (DCCEE), Canberra, ACT, Australia, 16 pp.
De’ath, G., J.M. Lough, and K.E. Fabricius, 2009: Declining coral calcification on the
Great Barrier Reef. Science, 323(5910), 116-119.
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.
Dean, S. and P. Stott, 2009: The Effect of local circulation variability on the detection
and attribution of New Zealand temperature trends. Journal of Climate, 22(23),
6217-6229.
Deo, R., J. Syktus, C. McAlpine, P. Lawrence, H. McGowan, and S. Phinn, 2009: Impact
of historical land cover change on daily indices of climate extremes including
droughts in eastern Australia. Geophysical Research Letters, 36(8), L08705,
doi:10.1029/2009GL037666.
Depczynski, M.R., J.P. Gilmour, T. Ridgway, H. Barnes, A.J. Heyward, T.H. Holmes, J.A.Y.
Moore, B.T. Radford, D. Thomson, P. Tinkler, and S.K. Wilson, 2013: Bleaching,
coral mortality and subsequent survivorship on a West Australian fringing reef.
Coral Reefs, 32, 233-238.
DERM, DIP, and LGAQ, 2010: Increasing Queensland’s Resilience to Inland Flooding
in a Changing Climate: Final Report on the Inland Flooding Study. Joint
report by Department of Environment and Resource Management (DERM),
Department of Infrastructure and Planning (DIP), and Local Government
Association of Queensland (LGAQ), State of Queensland, Brisbane, QLD,
Australia, 12 pp.
Derraik, J.G.B. and D. Slaney, 2007: Anthropogenic environmental change, mosquito-
borne diseases and human health in New Zealand. EcoHealth, 4(1), 72-81.
Derraik, J.G.B., D. Slaney, E.R. Nye, and P. Weinstein, 2010: Chikungunya virus: a
novel and potentially serious threat to New Zealand and the South Pacific
Islands. American Journal of Tropical Medicine and Hygiene, 83(4), 755-759.
Diamond, H.J., A.M. Lorrey, and J.A. Renwick, 2012: A southwest Pacific tropical
cyclone climatology and linkages to the El Niño–Southern Oscillation. Journal
of Climate, 26(1), 3-25.
Diaz-Pulido, G., L.J. McCook, S. Dove, R. Berkelmans, G. Roff, D.I. Kline, S. Weeks, R.D.
Evans, D.H. Williamson, and O. Hoegh-Guldberg, 2009: Doom and boom on a
resilient reef: climate change, algal overgrowth and coral recovery. PLoS ONE,
4(4), e5239, doi:10.1371/journal.pone.0005239.
DIICCSRTE, 2013: Climate Adaptation Outlook. A Proposed National Adaptation
Assessment Framework. Department of Industry, Innovation, Climate Change,
Science, Research and TertiaryEducation (DIICCSRTE), Canberra, ACT, Australia,
110 pp.
DNP, 2010: Kakadu National Park Climate Change Strategy 2010-2015. Director of
National Parks (DNP), Department of the Environment, Water, Heritage and the
Arts (DEWHA), Canberra, ACT, Australia, 15 pp.
Dobes, L., 2010: Notes on Applying ‘Real Options’ to Climate Change Adaptation
Measures, with Examples from Vietnam. CCEP working paper 7.10, Centre for
Climate Economics & Policy (CCEP), Crawford School of Economics and
Government, The Australian National University (ANU), Canberra, ACT, Australia,
23 pp.
Doherty, T.J. and S. Clayton, 2011: The psychological impacts of global climate
change. American Psychologist, 66(4), 265-276.
Donohue, R.J., T.R. McVicar, and M.L. Roderick, 2009: Climate-related trends in
Australian vegetation cover as inferred from satellite observations, 1981-2006.
Global Change Biology, 15(4), 1025-1039.
Dovers, S., 2009: Normalizing adaptation. Global Environmental Change, 19(1), 4-
6.
Dovers, S., 2013: How much adaptation: are existing policy and institutions enough?
In: Climate Adaptation Futures [Palutikof, J., S.L. Boulter, A.J. Ash, M. Stafford-
Smith, M. Parry, M. Waschka, and D. Guitart (eds.)]. John Wiley & Sons, Brisbane,
QLD, Australia, pp. 97-102.
Driscoll, D.A., D.B. Lindenmayer, A.F. Bennett, M. Bode, R.A. Bradstock, G.J. Cary, M.F.
Clarke, N. Dexter, R. Fensham, G. Friend, M. Gill, S. James, G. Kay, D.A. Keith, C.
MacGregor, J. Russell-Smith, D. Salt, J.E.M. Watson, R.J. Williams, and A. York,
2010: Fire management for biodiversity conservation: key research questions
and our capacity to answer them. Biological Conservation, 143(9), 1928-1939.
Drucker, A.G., G.P. Edwards, and W.K. Saalfeld, 2010: Economics of camel control in
central Australia. The Rangeland Journal, 32(1), 117-127.
DSE, 2007: Our Water Our Future: The Next Stage of the Government’s Water Plan.
Victorian Government Department of Sustainability and Environment (DSE),
Melbourne, VIC, Australia, 28 pp.
DSE, 2011: Guidelines for the Development of a Water Supply Demand Strategy.
Victorian Government Department of Sustainability and Environment (DSE),
Melbourne, VIC, Australia, 95 pp.
DSE, 2013: Victorian Climate Change Adaptation Plan. Victorian Government
Department of Sustainability and Environment (DSE), Melbourne, VIC, Australia,
96 pp.
DSEWPC, 2011: Regional Natural Resources Management Planning for Climate
Change Fund. Department of the Environment [formerly Department of
Sustainability, Environment, Water, Population and Communities (DSEWPC)],
Canberra, ACT, Australia, www.environment.gov.au/cleanenergyfuture/regional-
fund.
Duckett, P.E., P.D. Wilson, and A.J. Stow, 2013: Keeping up with the neighbours: using
a genetic measurement of dispersal and species distribution modelling to
assess impact of climate change on an Australian aid zone gecko (Gehyra
variegata). Diversity and Distributions, 19(8), 964-976.
Duncan, M. and G. Smart, 2011: 2
nd
Toolbox Tray: Assess the likely hazard: Bin 2.1
Flooding, Tool 2.1.4: Inundation modelling of present day and future floods. In:
Impacts of Climate Change on Urban Infrastructure and the Build Environment:
Toolbox Handbook. National Institute of Water and Atmospheric Research
(NIWA), MWH New Zealand Ltd., Branz, and GNS Science, NIWA, Wellington,
New Zealand, 13 pp. www.niwa.co.nz/sites/niwa.co.nz/files/tool_2.1.4_
innundation_modelling_0.pdf.
Dunlop, M. and P.R. Brown, 2008: Implications of Climate Change for Australia’s
National Reserve System: A Preliminary Assessment. Report to the Department
of Climate Change (DCC) and the Department of the Environment, Water,
Heritage and the Arts (DEWHA), DCC, Canberra, ACT, Australia, 147 pp.
Dunlop, M., D.W. Hilbert, S. Ferrier, A. House, A. Liedloff, S.M. Prober, A. Smyth, T.G.
Martin, T. Harwood, K.J. Williams, C. Fletcher, and H. Murphy, 2012: The
Implications of Climate Change for Biodiversity Conservation and the National
Reserve System: Final Synthesis. Report prepared for the Department of
Sustainability, Environment, Water, Population and Communities, and the
Department of Climate Change and Energy Efficiency by the Commonwealth
Scientific and Industrial Research Organization (CSIRO) Climate Adaptation
Flagship, Canberra, ACT, Australia, 79 pp.
Dunn, M.R., R.J. Hurst, J.A. Renwick, R.I.C.C. Francis, J. Devine, and A. McKenzie, 2009:
Fish Abundance and Climate Trends in New Zealand. New Zealand Aquatic
Environment and Biodiversity Report No. 31, Ministry of Fisheries, Wellington,
New Zealand, 73 pp.
Dupont, A., 2008: The strategic implications of climate change. Survival: Global
Politics and Strategy, 50(3), 29-54.
1420
Chapter 25 Australasia
25
Dupont, A. and G. Pearman, 2006: Heating Up the Planet: Climate Change and
Security. Lowy Institute Paper 12, Lowy Institute for International Policy, Sydney,
NSW, Australia, 143 pp.
Dwyer, A., C. Zoppou, O. Nielsen, S. Day, and S. Roberts, 2004: Quantifying Social
Vulnerability: A Methodology for Identifying Those at Risk to Natural Hazards.
Geoscience Australia Record 2004/14, Geoscience Australia, Canberra, ACT,
Australia, 92 pp.
ECAN, 2005: Regional Coastal Environment Plan for the Canterbury Region. Vol. 1:
incorporating plan changes 1, 2, and 4, and deleting all references to Restricted
Coastal Activities, Report No. R12/86, Environment Canterbury (ECAN),
Christchurch, New Zealand, 268 pp., ecan.govt.nz/publications/Plans/
RegionalCoastalEnvPlanNovember05.pdf.
Eckard, R.J. and B.R. Cullen, 2011: Impacts of future climate scenarios on nitrous
oxide emissions from pasture based dairy systems in south eastern Australia.
Animal Feed Science and Technology, 166-167, 736-748.
Edge Environment, 2011: Barriers to Effective Climate Change Adaptation:
Submissions. Edge Environment submission to Productivity Commission Issues
Paper, No. 54, Productivity Commission, Canberra, ACT, Australia, 7 pp.
Edwards, B. and M. Gray, 2009: A sunburnt country: the economic and financial
impact of drought on rural and regional families in Australia in an era of climate
change. Journal of Labour Economics, 12(1), 109-131.
Edwards, F., J. Dixon, G. Hall, K. Larsen, S. Lockie, B. Wood, M. Lawrence, I. Hanigan,
A. Hogan, and L. Hattersley, 2011: Climate change adaptation at the intersection
of food and health. Asia-Pacific Journal of Public Health, 23(2 Suppl.), S91-S104.
Elith, J., M. Kearney, and S. Phillips, 2010: The art of modelling range-shifting species.
Methods in Ecology and Evolution, 1(4), 330-342.
Ellemor, H., 2005: Reconsidering emergency management and indigenous communities
in Australia. Global Environmental Change Part B: Environmental Hazards, 6(1),
1-7.
Energy Futures Forum, 2006: The Heat Is On. The Future of Energy in Australia.
Commonwealth Scientific and Industrial Research Organization (CSIRO),
Canberra, ACT, Australia, 120 pp.
Evans, J.P. and I. Boyer-Souchet, 2012: Local sea surface temperatures add to extreme
precipitation in northeast Australia during La Niña. Geophysical Research
Letters, 39(10), L10803, doi:10.1029/2012GL052014.
Evans, L.S., P. Fidelman, C. Hicks, C. Morgan, A.L. Perry, and R. Tobin, 2011: Limits to
Climate Change Adaptation in the Great Barrier Reef: Scoping Ecological and
Social Limits. Final Report to the National Climate Change Adaptation Research
Facility (NCCARF), NCCARF, Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 65 pp.
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.
Farbotko, C. and H. Lazrus, 2012: The first climate refugees? Contesting global narratives
of climate change in Tuvalu. Global Environmental Change, 22(2), 382-390.
Fawcett, R.J.B., B.C. Trewin, K. Braganza, R.J. Smalley, B. Jovanovic, and D.A. Jones,
2012: On the Sensitivity of Australian Temperature Trends and Variability to
Analysis Methods and Observation Networks. CAWCR Technical Report No.050,
Centre for Australian Weather and Climate Research (CAWCR), A partnership
between the Bureau of Meteorology (BoM) and CSIRO, CAWCR, Melbourne,
VIC, Australia, 54 pp.
Feng, M., M.J. McPhaden, S.P. Xie, and J. Hafner, 2013: La Niña forces unprecedented
Leeuwin Current warming in 2011. Scientific Reports, 3, 1277, doi:10.1038/
srep01277.
Fensham, R.J., R.J. Fairfax, and D.P. Ward, 2009: Drought-induced tree death in
savanna. Global Change Biology, 15(2), 380-387.
Fidelman, P.I.J., A.M. Leitch, and D.R. Nelson, 2013: Unpacking multilevel adaptation
to climate change in the Great Barrier Reef, Australia. Global Environmental
Change, 23(4), 800-812.
Figueira, W.F. and D.J. Booth, 2010: Increasing ocean temperatures allow tropical
fishes to survive overwinter in temperate waters. Global Change Biology, 16(2),
506-516.
Figueira, W.F., P. Biro, D.J. Booth, and V.C. Valenzuela, 2009: Performance of tropical fish
recruiting to temperate habitats: role of ambient temperature and implications
of climate change. Marine Ecology Progress Series, 384, 231-239.
Finger, R., P. Lazzarotto, and P. Calanca, 2010: Bio-economic assessment of climate
change impacts on managed grassland production. Agricultural Systems,
103(9), 666-674.
Finlay, K.J., A.L. Yen, J.P. Aurambout, G.A.C. Beattie, P. Barkley, and J.E. Luck, 2009:
Consequences for Australian biodiversity with establishment of the asiatic citrus
psyllid, diaphorina citri, under present and future climates. Biodiversity, 10(2-3),
25-32.
Fitzgerald, C., 2009: Implementing residential water use targets a Melbourne
perspective. In: 5th IWA Specialist Conference on Efficient Use and Management
of Urban Water Supply, 25-28 October 2009. International Water Association
(IWA), Sydney, NSW, Australia, 7 pp.
Fitzgerald, G., R. Norton, M. Tausz, G. O’Leary, S. Seneweera, S. Posch, M. Mollah, J.
Brand, R. Armstrong, N. Mathers, J. Luck, W. Griffiths, and P. Trebicki, 2010: Future
effects of elevated CO
2
on wheat production an overview of FACE research in
Victoria, Australia. In: Proceedings of Food Security from Sustainable Agriculture,
15th Agronomy Conference, 15-18 November 2010, Lincoln, New Zealand
[Dove, H. and R.A. Culvenor (eds.)]. Australian Society of Agronomy, Gosford,
NSW, Australia, www.regional.org.au/au/asa/2010/climate-change/co2/6996_
fitzgeraldgj.htm#TopOfPage.
Fitzharris, B.B., 2004: Possible impact of future climate change on seasonal snow of
the Southern Alps of New Zealand. In: Glimpses of a Gaian World: Essays in
Honour of Peter Holland [Kearsley, G.W., B.B. Fitzharris, and P.G. Holland (ed.)].
Department of Geography, School of Social Science, University of Otago,
Dunedin, New Zealand, pp. 231-241.
Fitzharris, B.B., 2010: Climate Change Impacts on Dunedin. Dunedin City Council,
Dunedin, New Zealand, 58 pp.
Fitzpatrick, M.C., A.D. Gove, N.J. Sanders, and R.R. Dunn, 2008: Climate change,
plant migration, and range collapse in a global biodiversity hotspot: the Banksia
(Proteaceae) of Western Australia. Global Change Biology, 14(6), 1337-1352.
Fitzsimons, P., M. Bond, and S. Webber, 2010: Creating a Participatory Adaptive
Capacity Index for Climate Change Adaptation: Report of Engagement Process in
the South-West of Victoria. Department of Primary Industries, State Government
of Victoria, Tatura, VIC, Australia, 41 pp.
Fletcher, C.S., B.M. Taylor, A.N. Rambaldi, B.P. Harman, S. Heyenga, K.R. Ganegodage,
F. Lipkin, and R.R.J. McAllister, 2013: Costs and Coasts: An Empirical Assessment
of Physical and Institutional Climate Adaptation Pathways: Final Report.
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 53 pp.
FOA and MPI, 2012: New Zealand Plantation Forestry Industry: Facts and Figures
2011/2012. New Zealand Forest Owners Association Inc. (FOA) and Ministry
for Primary Industries (MPI), Wellington, New Zealand, 45 pp.
Fouquet, A., G.F. Ficetola, A. Haigh, and N. Gemmell, 2010: Using ecological niche
modelling to infer past, present and future environmental suitability for
Leiopelma hochstetteri, an endangered New Zealand native frog. Biological
Conservation, 143(6), 1375-1384.
Frame, B., M. Brignall-Theyer, R. Taylor, and K. Delaney, 2007: Four Future Scenarios
for New Zealand: Work in Progress Edition 2. Developed by the Landcare
Research Scenarios Working Group, Manaaki Whenua Press, Lincoln, New
Zealand, 108 pp.
Frederiksen, C.S., J.S. Frederiksen, J.M. Sisson, and S.L. Osbrough, 2011: Changes
and projections in Australian winter rainfall and circulation: anthropogenic forcing
and internal variablilty. International Journal of Climate Change Impacts and
Responses, 2, 143-162.
Frederiksen, J.S. and C.S. Frederiksen, 2007: Interdecadal changes in southern
hemisphere winter storm track modes. Tellus Series A, 59(5), 599-617.
Freeman, C. and C. Cheyne, 2008: Coasts for sale: gentrification in New Zealand.
Planning Theory and Practice, 9(1), 33-56.
Fritze, J.G., G.A. Blashki, S. Burke, and J. Wiseman, 2008: Hope, despair and
transformation: climate change and the promotion of mental health and
wellbeing. International Journal of Mental Health Systems, 2(1), 13, doi:10.1186/
1752-4458-2-13.
Fuentes, M.M.P.B., M. Hamann, and V. Lukoschek, 2009: Marine reptiles. In: A Marine
Climate Change Impacts and Adaptation Report Card for Australia 2009
[Poloczanska, E.S., A.J. Hobday, and A.J. Richardson (eds.)]. Publication 05/09,
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 18 pp., www.ocean
climatechange.org.au/content/images/uploads/MarineReptiles_FINAL.pdf.
Fulton, E.A., 2011: Interesting times: winners, losers, and system shifts under climate
change around Australia. Ices Journal of Marine Science, 68(6), 1329-1342.
Fünfgeld, H., B. Webb, and D. McEvoy, 2012: The significance of adaptation framing
in local and regional climate change adaptation initiatives in Australia. In:
Resilient Cities 2: Cities and Adaptation to Climate Change –Proceedings of
1421
Australasia Chapter 25
25
the Global Forum 2011 [Otto-Zimmermann, K. (ed.)]. Springer, Dordrecht,
Netherlands, pp. 283-294.
Funk, J. and S. Kerr, 2007: Restoring forests through carbon farming on Māori land
in New Zealand/Aotearoa. Mountain Research and Development, 27(3), 202-
205.
Gallagher, R.V., D.E. Duursma, J. O’Donnell, P.D. Wilson, P.O. Downey, L. Hughes, and
M.R. Leishman, 2012a: The grass may not always be greener: projected reductions
in climatic suitability for exotic grasses under future climates in Australia.
Biological Invasions, 15(5), 961-975.
Gallagher, R.V., L. Hughes, and M.R. Leishman, 2012b: Species loss and gain in
communities under future climate change: consequences for functional diversity.
Ecography, 36, 531-540.
Gallant, A.J.E. and D.J. Karoly, 2010: A combined climate extremes index for the
Australian region. Journal of Climate, 23, 6153-6165.
Gallant, A.J.E., K.J. Hennessy, and J. Risbey, 2007: Trends in rainfall indices for six
Australian regions: 1910-2005. Australian Meteorological Magazine, 56, 223-
239.
Gallant, A.J.E., A.S. Kiem, D.C. Verdon-Kidd, R.C. Stone, and D.J. Karoly, 2012:
Understanding hydroclimate processes in the Murray-Darling Basin for natural
resources management. Hydrology and Earth System Sciences, 16(7), 2049-
2068.
Gardiner, L., D. Firestone, G. Waibl, N. Mistal, K. Van Reenan, D. Hynes, J. Byfield, S.
Oldfield, S. Allan, B. Kouvelis, J. Smart, A. Tait, and A. Clark, 2009: Climate
Change Effects on the Land Transport Network Volume One: Literature Review
and Gap Analysis. NZ Transport Agency Research Report No 378, New Zealand
Transport Agency, Wellington, New Zealand, 226 pp.
Gardner, J., A.-M. Dowd, C. Mason, and P. Ashworth, 2009a: A Framework for
Stakeholder Engagement on Climate Adaptation. CSIRO Climate Adaptation
Flagship Working Paper No. 3, Commonwealth Scientific and Industrial Research
Organization (CSIRO), Kenmore, QLD, Australia, 31 pp.
Gardner, J.L., R. Heinsohn, and L. Joseph, 2009b: Shifting latitudinal clines in avian
body size correlate with global warming in Australian passerines. Proceedings
of the Royal Society B, 276(1674), 3845-3852.
Gardner, J., R. Parsons, and G. Paxton, 2010: Adaptation Benchmarking Survey:
Initial Report. CSIRO Climate Adaptation Flagship Working Paper No. 4,
Commonwealth Scientific and Industrial Research Organization (CSIRO),
Canberra, ACT, Australia, 58 pp.
Garnaut, R., 2008: The Garnaut Climate Change Review: Final Report. Cambridge
University Press, Melbourne, VIC, Australia, 634 pp.
Garnett, S., D. Franklin, G. Ehmke, J. VanDerWal, L. Hodgson, C. Pavey, A. Reside, J.
Welbergen, S. Butchart, G. Perkins, and S. Williams, 2013: Climate Change
Adaptation Strategies for Australian Birds. National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 109 pp.
Gaydon, R.S., H.G. Beecher, R. Reinke, S. Crimp, and S.M. Howden, 2010: Rice. In:
Adapting Agriculture to Climate Change. Preparing Australian Agriculture,
Forestry and Fisheries for the Future [Stokes, C.J. and S.M. Howden (eds.)].
Commonwealth Scientific and Industrial Research Organization (CSIRO)
Publishing, Collingwood, VIC, Australia, pp. 67-83.
GBRMPA, 2007: Great Barrier Reef Climate Change Action Plan 2007-2012. Great
Barrier Reef Marine Park Authority (GBRMPA), Commonwealth of Australia,
Townsville, QLD, Australia, 10 pp.
GBRMPA, 2009a: Great Barrier Reef Outlook Report 2009. Great Barrier Reef Marine
Park Authority (GBRMPA), Townsville, QLD, Australia, 192 pp.
GBRMPA, 2009b: Great Barrier Reef Tourism Climate Change Action Strategy 2009-
2012. Great Barrier Reef Marine Park Authority (GBRMPA), Townsville, QLD,
Australia, 52 pp.
GBRMPA, 2013: Great Barrier Reef Report Card 2011: Reef Water Quality Protection
Plan. Reef Water Quality Protection Plan Secretariat, Queensland Department
of the Premier and Cabinet, Brisbane, QLD, Australia, 6 pp.
Gerard, P.J., J.R.F. Barringer, J.G. Charles, S.V. Fowler, J.M. Kean, C.B. Phillips, A.B. Tait,
and G.P. Walker, 2012: Potential effects of climate change on biological control
systems: case studies from New Zealand. BioControl, 58(2), 149-162.
GETF, 2011: Roadmap to Renewable and Low Emission Energy in Remote Communities.
Report by the Green Energy Task Force (GETF), Northern Territory Government,
Darwin, NT, Australia, 65 pp.
Gibbs, M. and T. Hill, 2012: Coastal Climate Change Risk Legal and Policy Responses
in Australia. Department of Climate Change and Energy Efficiency, Canberra,
ACT, Australia, 89 pp.
Gibson, L., A. McNeill, P. de Tores, A. Wayne, and C. Yates, 2010: Will future climate
change threaten a range restricted endemic species, the quokka (Setonix
brachyurus), in south west Australia? Biological Conservation, 143(11), 2453-
2461.
Gifford, R., 2011: The dragons of inaction: psychological barriers that limit climate
change mitigation and adaptation. American Psychologist, 66(4), 290-302.
Gifford, R., L. Scannell, C. Kormos, L. Smolova, A. Biel, S. Boncu, V. Corral, H. Güntherf,
K. Hanyu, D. Hine, F.G. Kaiser, K. Korpela, L.M. Lima, A.G. Mertig, R.G. Mira, G.
Moser, P. Passafaro, J.Q. Pinheiro, S. Saini, T. Sako, E. Sautkina, Y. Savina, P.
Schmuck, W. Schultz, K. Sobeck, E.-L. Sundblad, and D. Uzzell, 2009: Temporal
pessimism and spatial optimism in environmental assessments: an 18-nation
study. Journal of Environmental Psychology, 29(1), 1-12.
Gillanders, B.M., T.S. Elsdon, I.A. Halliday, G.P. Jenkins, J.B. Robins, and F.J. Valesini,
2011: Potential effects of climate change on Australian estuaries and fish
utilising estuaries: a review. Marine and Freshwater Research, 62(9), 1115-
1131.
Giltrap, D., A.-G. Ausseil, J. Ekanayake, S.M. Pawson, P. Hall, P. Newsome, and J.
Dymond, 2009: Chapter 2: Environmental impacts of large-scale forestry for
bioenergy. In: Bioenergy Options for New Zealand. Analysis of Large-Scale
Bioenergy from Forestry [Hall, P. and M. Jack (eds.)]. Scion, Rorotua, New
Zealand, pp. 71-121.
Glavovic, B., W. Saunders, and J. Becker, 2010: Land-use planning for natural hazards
in New Zealand: the setting, barriers, ‘burning issuesand priority actions.
Natural Hazards, 54(3), 679-706.
Gorddard, R., R. Wise, K. Alexander, A. Langston, A. Leitch, M. Dunlop, A. Ryan, and
J. Langridge, 2012: Striking the Balance: Coastal Development and Ecosystem
Values. Department of Climate Change and Energy Efficiency, Canberra, ACT,
Australia, 107 pp.
Gordon, D.P., J. Beaumont, A. MacDiarmid, D.A. Robertson, and S.T. Ahyong, 2010:
Marine Biodiversity of Aotearoa New Zealand. PLoS ONE, 5(8), e10905,
doi:10.1371/journal.pone.0010905.
Gorman-Murray, A., 2008: Before and after climate change: the snow country in
Australian imaginaries. M/C Journal, 11(5), journal.media-culture.org.au/
index.php/mcjournal/article/viewArticle/65.
Gorman-Murray, A., 2010: An Australian feeling for snow: towards understanding
cultural and emotional dimensions of climate change. Cultural Studies Review,
16(1), 60-81.
Gosling, E. and K.J.H. Williams, 2010: Connectedness to nature, place attachment
and conservation behaviour: testing connectedness theory among farmers.
Journal of Environmental Psychology, 30(3), 298-304.
Gosling, S.N., G.R. McGregor, and J.A. Lowe, 2009: Climate change and heat-related
mortality in six cities. Part 2: climate model evaluation and projected impacts
from changes in the mean and variability of temperature with climate change.
International Journal of Biometeorology, 53(1), 31-51.
Goss, J. and B. Lindquist, 2000: Placing movers: an overview of the Asia-Pacific
migration system. Contemporary Pacific, 12(2), 385-414.
Grace, W.J., V.O. Sadras, and P.T. Hayman, 2009: Modelling heatwaves in viticultural
regions of southeastern Australia. Australian Meteorological and Oceanographic
Journal, 58(4), 249-262.
Grafton, R.Q. and K. Hussey, 2007: Buying back the living Murray: at what price?
Australasian Journal of Environmental Management, 14(2), 74-81.
Graham, P., L. Reedman, and F. Poldy, 2008: Modelling of the Future of Transport Fuels
in Australia. IR 1046, Report to the Future Fuels Forum by the Commonwealth
Scientific and Industrial Research Organization (CSIRO), Canberra, ACT,
Australia, 111 pp.
Gray, M.C., J.C. Altman, and N. Halasz, 2005: The Economic Value of Wild Resources
to the Indigenous Community of the Wallis Lake Catchment. CAEPR Discussion
Paper No 272/2005, Centre for Aboriginal Economic Policy Research (CAEPR),
The Australian National University (ANU), Canberra, ACT, Australia, 40 pp.
Gray, W., R. Ibbitt, R. Turner, M. Duncan, and M. Hollis, 2005: A Methodology to Assess
the Impacts of Climate Change on Flood Risk in New Zealand. National Institute
of Water and Atmospheric Research (NIWA) Client Report CHC2005-060,
Prepared for New Zealand Climate Change Office, Ministry for the Environment,
NIWA, Christchurch, New Zealand, 36 pp.
Green, A.E., M.R. Power, and D.M. Jang, 2008: Trans-Tasman migration: New Zealanders’
explanations for their move. New Zealand Geographer, 64(1), 34-45.
Green, B.S., C. Gardner, A. Linnane, and P.J. Hawthorne, 2010: The good, the bad and
the recovery in an assisted migration. PLoS ONE, 5(11), e14160, doi:10.1371/
journal.pone.0014160.
1422
Chapter 25 Australasia
25
Green, D., S. Jackson, and J. Morrison (eds.), 2009: Risks from Climate Change to
Indigenous Communities in the Tropical North of Australia. Department of
Climate Change and Energy Efficiency, Canberra, ACT, Australia, 185 pp.
Green, D., L. Alexander, K. McInnes, J. Church, N. Nicholls, and N. White, 2010a: An
assessment of climate change impacts and adaptation for the Torres Strait
Islands, Australia. Climatic Change, 102(3-4), 405-433.
Green, D., J. Billy, and A. Tapim, 2010b: Indigenous Australians’ knowledge of weather
and climate. Climatic Change, 100(2), 337-354.
Green, K., J.A. Stein, and M.M. Driessen, 2008: The projected distributions of
Mastacomys fuscus and Rattus lutreolus in south-eastern Australia under a
scenario of climate change: potential for increased competition? Wildlife
Research, 35(2), 113-119.
Griffiths, G.M., 2007: Changes in New Zealand daily rainfall extremes 1930-2004.
Weather and Climate, 27, 3-44.
Griffiths, G.M., L.E. Chambers, M.R. Haylock, M.J. Manton, N. Nicholls, H.J. Baek, Y.
Choi, P.M. Della-Marta, A. Gosai, N. Iga, R. Lata, V. Laurent, L. Maitrepierre, H.
Nakamigawa, N. Ouprasitwong, D. Solofa, L. Tahani, D.T. Thuy, L. Tibig, B. Trewin,
K. Vediapan, and P. Zhai, 2005: Change in mean temperature as a predictor of
extreme temperature change in the Asia-Pacific region. International Journal
of Climatology, 25(10), 1301-1330.
Guillaume, J., G.M. Li, M.F. Hutchinson, K. Proust, and S. Dovers, 2010: A Differential
Vulnerability Assessment of Climate Change Impacts on Darwin. Integrated
Assessment of Climate Change Impacts on Urban Settlements (IACCIUS) Project,
Fenner School of Environment and Society, The Australian National University,
Canberra, ACT, Australia, 173 pp.
Gunasekera, D., Y. Kim, C. Tulloh, and M. Ford, 2007: Climate change: impacts on
Australian agriculture. Australian Commodities, 14, 657-676.
Gunasekera, D., M. Ford, E. Heyhoe, A. Gurney, H. Ahammad, S.J. Phipps, I.N. Harman,
J.J. Finnigan, and M. Brede, 2008: Global integrated assessment model: a new
analytical tool for assessing climate change risks and policies. Australian
Commodities, 15(1), 195-216.
Gurran, N., 2008: The turning tide: amenity migration in coastal Australia.
International Planning Studies, 13(4), 391-414.
Gurran, N., C. Squires, and E.J. Blakely, 2006: Meeting the Sea Change Challenge: Best
Practice Models of Local & Regional Planning for Sea Change Communities.
Report No. 2 for the National Sea Change Taskforce, The University of Sydney
Planning Research Centre, Sydney, NSW, Australia, 130 pp.
Gurran, N., E. Hamin, and B. Norman, 2008: Planning for Climate Change: Leading
Practice Principles and Models for Sea Change Communities in Coastal
Australia. Prepared for the National Sea Change Taskforce, The University of
Sydney, Faculty of Architecture, Sydney, NSW, Australia, 66 pp.
Hagger, V., D. Fisher, S. Schmidt, and S. Blomberg, 2013: Assessing the vulnerability
of an assemblage of subtropical rainforest vertebrate species to cliamte change
in south-east Queensland. Austral Ecology, 38, 465-475.
Hall, A. and G.V. Jones, 2009: Effect of potential atmospheric warming on temperature-
based indices describing Australian winegrape growing conditions. Australian
Journal of Grape and Wine Research, 15(2), 97-119.
Hallegraeff, G.M., 2010: Ocean climate change, phytoplankton community responses,
and harmful algal blooms: a formidable predictive challenge. Journal of
Phycology, 46(2), 220-235.
Hamin, E.M. and N. Gurran, 2009: Urban form and climate change: balancing
adaptation and mitigation in the US and Australia. Habitat International, 33(3),
238-245.
Handmer, J. and K. Haynes, 2008: Community Bushfire Safety. Commonwealth
Scientific and Industrial Research Organization (CSIRO) Publishing, Collingwood,
VIC, Australia, 208 pp.
Hanigan, I.C., C.D. Butler, P.N. Kokic, and M.F. Hutchinson, 2012: Suicide and drought
in New South Wales, Australia, 1970-2007. Proceedings of the National
Academy of Sciences of the United States of America, 109(35), 13950-13955.
Hanna, E.G., T. Kjellstrom, C. Bennett, and K. Dear, 2011: Climate change and rising
heat: population health implications for working people in Australia. Asia
Pacific Journal of Public Health, 23(Suppl. 2), S14-S26.
Hannah, J., 2004: An updated analysis of long-term sea level change in New Zealand.
Geophysical Research Letters, 31(3), L03307, doi:10.1029/2003GL019166.
Hannah, J. and R. Bell, 2012: Regional sea level trends in New Zealand. Journal of
Geophysical Research: Oceans, 117, C01004, doi:10.1029/2011JC007591.
Hansen, A., P. Bi, M. Nitschke, P. Ryan, D. Pisaniello, and G. Tucker, 2008: The effect of
heat waves on mental health in a temperate Australian city. Environmental
Health Perspectives, 116(10), 1369-1375.
Hanslow, K., D. Gunasekera, B.R. Cullen, and D. Newth, 2013: Economic impacts of
climate change on the Australian dairy sector. Australian Journal of Agricultural
and Resource Economics, 58(1), 60-77.
Harley, D., P. Bi, G. Hall, A. Swaminathan, S.L. Tong, and C. Williams, 2011: Climate
change and infectious diseases in Australia: future prospects, adaptation options,
and research priorities. Asia-Pacific Journal of Public Health, 23(2 Suppl.), 54-66.
Harman, I.N., M. Ford, G. Jakeman, S.J. Phipps, M. Brede, J.J. Finnigan, D. Gunasekera,
and H. Ahammad, 2008: Assessment of Future Global Scenarios for the Garnaut
Climate Change Review: An Application of the GIAM Framework. Commonwealth
Scientific and Industrial Research Organization (CSIRO), Canberra, ACT,
Australia, 51 pp.
Harmsworth, G., 2003: Māori Perspectives on Kyoto Policy: Interim Results. Reducing
Greenhouse Gas Emissions from the Terrestrial Biosphere (C09X0212), Landcare
Research NZ Ltd., Palmerston North, New Zealand, 31 pp.
Harmsworth, G. and B. Raynor, 2005: Cultural consideration in landslide risk
perception. In: Landslide Hazard and Risk [Glade, T., M.G. Anderson, and M.J.
Crozier (eds.)]. John Wiley & Sons, London, UK, pp. 219-249.
Harmsworth, G., M. Tahi, and C.K. Insley, 2010: Climate Change Business Opportunities
for Māori Land and Māori Organisations: Sustainable Land Management
Mitigation and Adaptation to Climate Change (SLMACC). Prepared for Ministry
of Agriculture and Forestry by Landcare Research New Zealand Ltd., Massey
University, Manaaki Whenua Press, Palmerston North, New Zealand, 59 pp.
Harper, B., T. Hardy, L. Mason, and R. Fryar, 2009: Developments in storm tide
modelling and risk assessment in the Australian region. Natural Hazards, 51(1),
225-238.
Harris, J.B.C., D.A. Fordham, P.A. Mooney, L.P. Pedler, M.B. Araujo, D.C. Paton, M.G.
Stead, M.J. Watts, H.R. Akcakaya, and B.W. Brook, 2012: Managing the long-
term persistence of a rare cockatoo under climate change. Journal of Applied
Ecology, 49(4), 785-794.
Harris, R.J. and G. Barker, 2007: Relative risk of invasive ants (Hymenoptera:
Formicidae) establishing in New Zealand. New Zealand Journal of Zoology,
34(3), 161-178.
Harsch, M.A., R. Buxton, R.P. Duncan, P.E. Hulme, P. Wardle, and J. Wilmshurst, 2012:
Causes of tree line stability: stem growth, recruitment and mortality rates over
15 years at New Zealand Nothofagus tree lines. Journal of Biogeography,
39(11), 2061-2071.
Hartog, J.R., A.J. Hobday, R. Matear, and M. Feng, 2011: Habitat overlap between
southern bluefin tuna and yellowfin tuna in the east coast longline fishery –
implications for present and future spatial management. Deep-Sea Research
Part II: Topical Studies in Oceanography, 58(5), 746-752.
Hassim, M.E.E. and K.J.E. Walsh, 2008: Tropical cyclone trends in the Australian re-
gion. Geochemistry, Geophysics, Geosystems, 9(7), Q07V07, doi:10.1029/
2007GC001804.
Hasson, A.E.A., G.A. Mills, B. Timbal, and K. Walsh, 2009: Assessing the impact of
climate change on extreme fire weather events over southeastern Australia.
Climate Research, 39(2), 159-172.
Hayward, B., 2008a: ‘Nowhere Far From the Sea’: political challenges of coastal
adaptation to climate change in New Zealand. Political Science, 60(1), 47-59.
Hayward, B., 2008b: Let’s talk about the weather: decentering democratic debate
about climate change. Hypathia, 23(3), 79-98.
Hayward, J.A., P.W. Graham, and P.K. Campbell, 2011: Projections of the Future Costs
of Electricity Generation Technologies: An Application of CSIRO’s Global and
Local Learning Model (GALLM). EP104982, Commonwealth Scientific and
Industrial Research Organization (CSIRO), Canberra, ACT, Australia, 78 pp.
HCCREMS, 2010: Potential Impacts of Climate Change on the Hunter, Central and
Lower North Coast of NSW. Hunter and Central Coast Regional Environmetal
Management Strategy (HCCREMS), Hunter Councils, Thornton, NSW, Australia,
86 pp.
HCCREMS, 2012: Decision Support for Coastal Adaptation: The Handbook. Hunter
and Central Coast Regional Environmetal Management Strategy (HCCREMS),
Hunter Councils, Thornton, NSW, Australia, 165 pp.
HDC, 2012: Long Term Plan 2012/22. Hastings District Council (HDC), Hastings, New
Zealand, 80 pp.
Heckbert, S., J. Russell-Smith, A. Reeson, J. Davies, G. James, and C. Meyer, 2012:
Spatially explicit benefit cost analysis of fire management for greenhouse
gas abatement. Austral Ecology, 37(6), 724-732.
Hellmann, J.J., J.E. Byers, B.G. Bierwagen, and J.S. Dukes, 2008: Five potential
consequences of climate change for invasive species. Conservation Biology,
22(3), 534-543.
1423
Australasia Chapter 25
25
Hendon, H.H., D.W.J. Thompson, and M.C. Wheeler, 2007: Australian rainfall and
surface temperature variations associated with the Southern Hemisphere
annular mode. Journal of Climate, 20(11), 2452-2467.
Hendrikx, J. and E. Hreinsson, 2012: The potential impact of climate change on
seasonal snow in New Zealand: part II industry vulnerability and future
snowmaking potential. Theoretical and Applied Climatology, 110(4), 619-630.
Hendrikx, J., E. Hreinsson, M. Clark, and B. Mullan, 2012: The potential impact of
climate change on seasonal snow in New Zealand: part I an analysis using
12 GCMs. Theoretical and Applied Climatology, 110(4), 607-618.
Hendrikx, J., C. Zammit, E. Hreinsson, and S. Becken, 2013: A comparative assessment
of the potential impact of climate change on the ski industry in New Zealand
and Australia. Climatic Change, 119(3-4), 965-978.
Hennessy, K.J., B. Fitzharris, B. Bates, N. Harvey, M. Howden, L. Hughes, J. Salinger,
and R. Warrick, 2007: Australia and New Zealand. In: Climate Change 2007:
Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the
Fourth Assessment Report of the Intergovernmental Panel on Climate Change
[Parry, M., O. Canziani, J. Palutikof, P. van der Linden, and C. Hanson, (eds.)].
Cambridge University Press, Cambridge, UK and New York, NY, USA, pp. 507-540.
Hennessy, K.J., R. Fawcett, D.G.C. Kirono, F.S. Mpelasoka, D. Jones, J.M. Bathols, P.H.
Whetton, M. Stafford Smith, M. Howden, C.D. Mitchell, and N. Plummer, 2008a:
An Assessment of the Impact of Climate Change on the Nature and Frequency
of Exceptional Climatic Events. Commonwealth Scientific and Industrial
Research Organization (CSIRO) Climate Adaptation Flagship, Aspendale, VIC,
Australia and Bureau of Meteorology (BoM), Melbourne, VIC, Australia, 33 pp.
Hennessy, K.J., P.H. Whetton, K. Walsh, I.N. Smith, J.M. Bathols, M. Hutchinson, and
J. Sharples, 2008b: Climate change effects on snow conditions in mainland
Australia and adaptation at ski resorts through snowmaking. Climate Research,
35(3), 255-270.
Hertzler, G., 2007: Adapting to climate change and managing climate risks by using
real options. Australian Journal of Agricultural Research, 58(10), 985-992.
Hine, D.W., W.J. Phillips, J.P. Reser, R.W. Cooksey, A.D.G. Marks, P.D. Nunn, S.E. Watt,
and M.C. Ellul, 2013: Enhancing Climate Change Communication: Strategies
for Profiling and Targeting Australian Interpretive Communities. Synthesis and
Integrative Research, Final Report, National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 86 pp.
Hitchings, T.R., 2009: Leptophlebiidae (Ephemeroptera) of the alpine region of the
Southern Alps, New Zealand. Aquatic Insects, 31(Suppl. 1), 595-601.
Hobday, A.J., 2010: Ensemble analysis of the future distribution of large pelagic
fishes off Australia. Progress in Oceanography, 86(1-2), 291-301.
Hobday, A.J., 2011: Sliding baselines and shuffling species: implications of climate
change for marine conservation. Marine Ecology, 32(3), 392-403.
Hobday, A.J. and E.S. Poloczanska, 2010: Fisheries and aquaculture. In: Adapting
Agriculture to Climate Change: Preparing Australian Agriculture, Forestry and
Fisheries for the Future [Stokes, C. and S. Howden (eds.)]. Commonwealth
Scientific and Industrial Research Organization (CSIRO) Publishing, Collingwood,
VIC, Australia, pp. 205-228.
Hobday, A.J., E.S. Poloczanska, and R. Matear (eds.), 2008: Implications of Climate
Change for Australian Fisheries and Aquaculture: A Preliminary Assessment.
Report to the Department of Climate Change, Canberra, ACT, Australia, 74 pp.
Hobson, K. and S. Niemeyer, 2011: Public responses to climate change: the role of
deliberation in building capacity for adaptive action. Global Environmental
Change, 21(3), 957-971.
Hodgkinson, J.H., A. Littleboy, M. Howden, K. Moffat, and B. Loechel, 2010a: Climate
Adaptation in the Australian Mining and Exploration Industries. CSIRO Climate
Adaptation Flagship Working Paper No. 5, Commonwealth Scientific and
Industrial Research Organization (CSIRO), Canberra, ACT, Australia, 27 pp.
Hodgkinson, J.H., B. Loechel, K. Moffat, M. Howden, A. Littleboy, and S. Crimp, 2010b:
Climate risks to the Australian mining industry a preliminary review of
vulnerabilities. In: Sustainable Mining 2010 Conference: 17-19 August 2010,
Kalgoorlie, Western Australia. The Australasian Institute of Mining and
Metallurgy, Carlton South, VIC, Australia, pp. 341-350.
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., 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.
Hoelzle, M., T. Chinn, D. Stumm, F. Paul, M. Zemp, and W. Haeberli, 2007: The
application of glacier inventory data for estimating past climate change effects
on mountain glaciers: a comparison between the European Alps and the Southern
Alps of New Zealand. Global and Planetary Change, 56(1-2), 69-82.
Hofmeester, C., B. Bishop, L. Stocker, and G. Syme, 2012: Social cultural influences
on current and future coastal governance. Futures, 44(8), 719-729.
Hogan, A., H.L. Berry, S.P. Ng, and A. Bode, 2011a: Decisions Made by Farmers Related
to Climate Change. Publication No. 10/28, Project No. PRJ-004546, Rural Industries
Research and Development Corporation, Barton, ACT, Australia, 64 pp.
Hogan, A., A. Bode, and H. Berry, 2011b: Farmer health and adaptive capacity in the
face of climate change and variability. Part 2: contexts, personal attributes and
behaviors. International Journal of Environmental Research and Public Health,
8(10), 4055-4068.
Holbrook, N.J., J. Davidson, M. Feng, A.J. Hobday, J.M. Lough, S. McGregor, and J.S.
Risbey, 2009: El Niño-Southern Oscillation. In: A Marine Climate Change Impacts
and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., A.J. Hobday,
and A.J. Richardson (eds.)]. Publication 05/09, National Climate Change
Adaptation Research Facility (NCCARF), Griffith University, Gold Coast Campus,
Southport, QLD, Australia, 25 pp., www.oceanclimatechange.org.au/content/
images/uploads/ENSO-final.pdf.
Hope, P., W. Drosdowsky, and N. Nicholls, 2006: Shifts in the synoptic systems
influencing southwest Western Australia. Climate Dynamics, 26(7-8), 751-764.
Hope, P., B. Timbal, and R. Fawcett, 2010: Associations between rainfall variability in
the southwest and southeast of Australia and their evolution through time.
International Journal of Climatology, 30(9), 1360-1371.
Hopkins, D., J.E.S. Higham, and S. Becken, 2012: Climate change in a regional context:
relative vulnerability in the Australasian skier market. Regional Environmental
Change, 13(2), 449-458.
Hovenden, M.J. and A.L. Williams, 2010: The impacts of rising CO
2
concentrations
on Australian terrestrial species and ecosystems. Austral Ecology, 35(6), 665-
684.
Howard, W.R., M. Nash, K. Anthony, K. Schmutter, H. Bostock, D. Bromhead, M. Byrne,
K. Currie, G. Diaz Pulido, S. Eggins, M. Ellwood, B. Eyre, R. Haese, G. Hallegraeff,
K. Hill, C. Hurd, C. Law, A. Lenton, R. Matear, B. McNeil, M. McCulloch, M.N.
Müller, P. Munday, B. Opdyke, J.M. Pandolfi, R. Richards, D. Roberts, B.D. Russell,
A.M. Smith, B. Tilbrook, A. Waite, and J. Williamson, 2012: Ocean acidification.
In: A Marine Climate Change Impacts and Adaptation Report Card for Australia
2012 [Poloczanska, E.S., A.J. Hobday, and A.J. Richardson (eds.)]. National
Climate Change Adaptation Research Facility (NCCARF), Gold Coast, QLD,
Australia, pp. 114-150, www.oceanclimatechange.org.au/content/images/
uploads/2012_OceanAcidification_MarineReportCard.pdf.
Howden, S.M. and S.J. Crimp, 2008: Drought and high temperature extremes: effects
on water and electricity demands. In: Transitions: Pathways Towards Sustainable
Urban Development in Australia [Newton, P.W. (ed.)]. Commonwealth Scientific
and Industrial Research Organization (CSIRO), Collingwood, VIC, Australia, pp.
227-244.
Howden, S.M. and C.J. Stokes, 2010: Introduction. In: Adapting Agriculture to Climate
Change: Preparing Australian Agriculture, Forestry and Fisheries for the Future
[Stokes, C.J. and S.M. Howden (eds.)]. Commonwealth Scientific and Industrial
Research Organization (CSIRO) Publishing, Collingwood, VIC, Australia, pp. 1-11.
Howden, S.M., S.J. Crimp, and C.J. Stokes, 2008: Climate change and Australian
livestock systems: impacts, research and policy issues. Australian Journal of
Experimental Agriculture, 48(7), 780-788.
Howden, S.M., S.J. Crimp, and R. Nelson, 2010: Australian agriculture in a climate of
change. In: Managing Climate Change: Papers from the GREENHOUSE 2009
Conference [Jubb, I., P. Holper, and W. Cai (eds.)]. Commonwealth Scientific and
Industrial Research Organization (CSIRO) Publishing, Collingwood, VIC,
Australia, pp. 101-111.
Howden-Chapman, P., 2010: Climate change & human health: impact and adaptation
issues for New Zealand. In: Climate Change Adaptation in New Zealand: Future
Scenarios and Some Sectoral Perspectives [Nottage, R., D. Wratt, J. Bornman,
and K. Jones (eds.)]. New Zealand Climate Change Centre, Wellington, New
Zealand, pp. 112-121.
Howe, C., R.N. Jones, S. Maheepala, and B. Rhodes, 2005: Melbourne Water Climate
Change Study: Implications of Potential Climate Change for Melbourne’s Water
Resources. Report CMIT-2005-104, Produced by Commonwealth Scientific and
Industrial Research Organization (CSIRO) Urban Water and CSIRO Atmospheric
Research in collaboration with Melbourne Water, Aspendale, VIC, Australia,
26 pp.
1424
Chapter 25 Australasia
25
Howells, E.J., R. Berkelmans, M.J.H. van Oppen, B.L. Willis, and L.K. Bay, 2013:
Historical thermal regimes define limits to coral acclimatization. Ecology, 94(5),
1078-1088.
HRC, 2010: Chapter 10: Natural hazards. In: Proposed One Plan. Horizons Regional
Council (HRC), Palmerston North, New Zealand, 8 pp., www.horizons.govt.nz/
assets/one-plan-august-2010/HRC_OP_Vol4_Chapter10.pdf.
Huang, C., A.G. Barnett, X. Wang, and S. Tong, 2012: The impact of temperature on
years of life lost in Brisbane, Australia. Nature Climate Change, 2(4), 265-270.
Hughes, J.D., K.C. Petrone, and R.P. Silberstein, 2012: Drought, groundwater storage
and streamflow decline in southwestern Australia. Geophysical Research
Letters, 39, L03408, doi:10.1029/2011GL050797.
Hughes, L., 2003: Climate change and Australia: trends, projections and impacts.
Austral Ecology, 28, 423-443.
Hughes, L., R. Hobbs, A. Hopkins, J. McDonald, M. Stafford-Smith, W. Steffen, and
S.E. Williams, 2010: National Climate Change Adaptation Research Plan:
Terrestrial Biodiversity. National Climate Change Adaptation Research Facility
(NCCARF), Griffith University, Gold Coast Campus, Southport, QLD, Australia,
70 pp.
Hughes, R. and D. Mercer, 2009: Planning to reduce risk: the Wildfire Management
Overlay in Victoria, Australia. Geographical Research, 47(2), 124-141.
Hughes, T.P., M.J. Rodrigues, D.R. Bellwood, D. Ceccarelli, O. Hoegh-Guldberg, L.
McCook, N. Moltschaniwskyj, M.S. Pratchett, R.S. Steneck, and B. Willis, 2007:
Phase shifts, herbivory, and the resilience of coral reefs to climate change.
Current Biology, 17(4), 360-365.
Hugo, G., 2010: Climate change-induced mobility and the existing migration regime
in Asia and the Pacific. In: Climate Change and Displacement [McAdam, J. (ed.)].
Hart Publishing, Oxford, UK, pp. 9-36.
Hugo, G., 2012: Population Distribution, Migration and Climate Change in Australia:
An Exploration. ACCARNSI Discussion Paper Node 2, Urban Management,
Transport and Social Inclusion, Australian Climate Change Adaptation Research
Network for Settlements and Infrastructure (ACCARNSI), one of eight networks
within the National Climate Change Adaptation Research Facility (NCCARF),
Griffith University, Gold Coast Campus, Southport, QLD, Australia, 93 pp.
Hunt, J., J. Altman, and K. May, 2009: Social Benefits of Aboriginal Engagement In
Natural Resource Management. CAEPR Working Paper 60, Centre for Aboriginal
Economic Policy Research (CAEPR), The Australian National University (ANU),
Canberra, ACT, Australia, 88 pp.
Hunter, E., 2009: ‘Radical hope’ and rain: climate change and the mental health of
Indigenous residents of northern Australia. Australasian Psychiatry, 17(6), 445-
452.
Hurlimann, A. and S. Dolnicar, 2011: Voluntary relocation an exploration of
Australian attitudes in the context of drought, recycled and desalinated water.
Global Environmental Change, 21(3), 1084-1094.
Hussey, K. and S. Dovers (eds.), 2007: Managing Water for Australia: The Social and
Institutional Challenges. Commonwealth Scientific and Industrial Research
Organization (CSIRO) Publishing, Collingwood, VIC, Australia, 157 pp.
IAA, 2011a: Barriers to Effective Climate Change Adaptation: Submissions. Institute
of Actuaries of Australia (IAA) submission to Productivity Commission Issues
Paper, No. 43, Productivity Commission, Canberra, ACT, Australia, 60 pp.,
www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions.
IAA, 2011b: Institute of Actuaries of Australia Submission to the Garnaut Climate
Change Review Update 2011. Institute of Actuaries of Australia (IAA), Sydney,
NSW, Australia, 28 pp., www.garnautreview.org.au/update-2011/commissioned-
work.html.
IAG, 2011: Barriers to Effective Climate Change Adaptation: Submissions. Insurance
Australia Group (IAG) submission to Productivity Commission Issues Paper, No.
39, Productivity Commission, Canberra, ACT, Australia, 24 pp., www.pc.gov.au/
projects/inquiry/climate-change-adaptation/submissions.
ICA, 2012: Historical Disaster Statistics., Insurance Council of Australia (ICA), Sydney,
NSW, Australia, online database, accessed 31 August 2013 at: www.insurance
council.com.au/industry-statistics-data/disaster-statistics/historical-disaster-
statistics.
ICNZ, 2013: Historical Events. Insurance Council of New Zealand (ICNZ), Wellington,
New Zealand, online database, accessed 31 August 2013, at www.icnz.org.nz/
natural-disaster/historic-events.
Insley, C.K., 2007: Māori Impacts from Emissions Trading Scheme: Interim High Level
Findings. 37 Degrees South Aotearoa, Gisborne, New Zealand, 22 pp.
Insley, C.K., 2010: Survey of Māori Business: Climate Change Māori Business
Opportunities. 37 Degrees South Aotearoa, Gisborne, New Zealand, 35 pp.
Insley, C.K. and R. Meade, 2008: Māori Impacts from the Emissions Trading Scheme:
Detailed Analysis and Conclusions. Prepared by 37 Degrees South Ltd. and
Cognitus Advisory Services Ltd. for the Ministry for the Environment, Wellington,
New Zealand, 48 pp.
IOCI, 2012: Western Australia’s Weather and Climate: A Synthesis of Indian Ocean
Climate Initiative Stage 3 Research: Summary for Policymakers [Bates, B., C.
Frederiksen, and J. Wormworth (eds.)]. Indian Ocean Climate Initiative (IIOCI),
Commonwealth Scientific and Industrial Research Organization (CSIRO) and
Bureau of Meteorology (BoM), BoM, Melbourne, VIC, Australia, 43 pp.
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.
Irving, A.D., S.D. Connell, and B.D. Russell, 2011: Restoring coastal plants to improve
global carbon storage: reaping what we sow. PLoS ONE, 6(3), e18311,
doi:10.1371/journal.pone.0018311.
Irving, D., P. Whetton, and A. Moise, 2012: Climate projections for Australia: a first
glance at CMIP5. Australian Meteorological and Oceanographic Journal, 62(4),
211-225.
Ishak, E.H., A. Rahman, S. Westra, A. Sharma, and G. Kuczera, 2013: Evaluating the
non-stationarity of Australian annual maximum flood. Journal of Hydrology,
494, 134-145.
Jackson, A.C., M.G. Chapman, and A.J. Underwood, 2008: Ecological interactions in
the provision of habitat by urban development: whelks and engineering by
oysters on artificial seawalls. Austral Ecology, 33(3), 307-316.
Jakes, P.J. and E.R. Langer, 2012: The adaptive capacity of New Zealand communities
to wildfire. International Journal of Wildland Fire, 21(6), 764-772.
Jakes, P.J., L. Kelly, and E.R. Langer, 2010: An exploration of a fire-affected community
undergoing change in New Zealand. Australian Journal of Emergency
Management, 25(3), 48-53.
Jakob, D., D. Karoly, and A. Seed, 2011: Non-stationarity in daily and sub-daily intense
rainfall – Part 2: regional assessment for sites in south-east Australia. Natural
Hazards and Earth System Sciences, 11(8), 2273-2284.
Jellyman, D.J., D.J. Booker, and E. Watene, 2009: Recruitment of Anguilla spp. glass
eels in the Waikato River, New Zealand. Evidence of declining migrations?
Journal of Fish Biology, 74(9), 2014-2033.
Jenkins, G.P., S.D. Conron, and A.K. Morison, 2010: Highly variable recruitment in an
estuarine fish is determined by salinity stratification and freshwater flow:
implications of a changing climate. Marine Ecology Progress Series, 417, 249-
261.
Jenkins, K.M., R.T. Kingsford, G.P. Closs, B.J. Wolfenden, C.D. Matthaei, and S.E. Hay,
2011: Climate change and freshwater systems in Oceania: an assessment of
vulnerability and adaptation opportunities. Pacific Conservation Biology, 17(3),
201-219.
Johnson, C.R., S.C. Banks, N.S. Barrett, F. Cazassus, P.K. Dunstan, G.J. Edgar, S.D. Frusher,
C. Gardner, M. Haddon, F. Helidoniotis, K.L. Hill, N.J. Holbrook, G.W. Hosie, P.R.
Last, S.D. Ling, J. Melbourne-Thomas, K. Miller, G.T. Pecl, A.J. Richardson, K.R.
Ridgway, S.R. Rintoul, D.A. Ritz, D.J. Ross, J.C. Sanderson, S.A. Shepherd, A.
Slotvvinski, K.M. Swadling, and N. Taw, 2011: Climate change cascades: shifts
in oceanography, species’ ranges and subtidal marine community dynamics
in eastern Tasmania. Journal of Experimental Marine Biology and Ecology,
400(1-2), 17-32.
Johnston, F.H., D.M. Kavanagh, D.M. Bowman, and R.K. Scott, 2002: Exposure to
bushfire smoke and asthma: an ecological study. Medical Journal of Australia,
176, 535-538.
Johnston, G., D. Burton, and M. Baker-Jones, 2013: Climate Change Adaptation in
the Boardroom: Combined Report. National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 85 pp.
Jollands, N., M. Ruth, C. Bernier, and N. Golubiewski, 2007: The climate’s long-term
impact on New Zealand infrastructure (CLINZI) project a case study of Hamilton
City, New Zealand. Journal of Environmental Management, 83(4), 460-477.
Jones, D., W. Wang, and R. Fawcett, 2009: High-quality spatial climate data-sets for
Australia. Australian Meteorological and Oceanographic Journal, 58(4), 233-248.
Jones, R.N., F. Chiew, W. Boughton, and L. Zhang, 2006: Estimating the sensitivity of
mean annual runoff to climate change using selected hydrological models.
Advances in Water Resources, 29(10), 1419-1429.
1425
Australasia Chapter 25
25
Jones, R., A. Wardell-Johnson, M. Gibberd, A. Pilgrim, G. Wardell-Johnson, J. Galbreath,
S. Bizjak, D. Ward, K. Benjamin, and J. Carlsen, 2010: The Impact of Climate
Change on the Margaret River Wine Region: Developing Adaptation and
Response Strategies for the Tourism Industry. Technical report by Sustainable
Tourism Cooperative Research Centre (STCRC), CRC for Sustainable Tourism Pty
Ltd., Griffith University, Gold Coast Campus, Southport, QLD, Australia, 84 pp.
Kamber, G., C. McDonald, and G. Price, 2013: Drying Out: Investigating the Economic
Effects of Drought in New Zealand. Reserve Bank of New Zealand Analytical
Note Series No. AN2013/02, Reserve Bank of New Zealand, Wellington, New
Zealand, 31 pp.
Kamman, C., L. Grünhage, U. Grüters, S. Janze, and H.-J. Jäger, 2005: Response of
aboveground grassland biomass and soil moisture to moderate long-term CO
2
enrichment. Basic and Applied Ecology, 6, 351-365.
Karoly, D.J. and S. Boulter, 2013: Afterword: floods, storms, fire and pestilence
disaster risk in Australia during 2010/11. In: Natural Disasters and Adaptation
to Climate Change [Boulter, S., J. Palutikof, D.J. Karoly, and D. Guitart (eds.)].
Cambridge University Press, Cambridge, UK, pp. 412-418.
Karoly, D.J. and K. Braganza, 2005: Attribution of recent temperature changes in the
Australian region. Journal of Climate, 18, 457-464.
Kawaguchi, S., A. Ishida, R. King, B. Raymond, N. Waller, A. Constable, S. Nicol, M.
Wakita, and A. Ishimatsu, 2013: Risk maps for Antarctic krill under projected
Southern Ocean acidification. Nature Climate Change, 3, 843-847.
Kay, R., A. Travers, and L. Dalton, 2013: The evolution of coastal vulnerability
assessments to support adaptive decision making in Australia: a review. In:
Climate Change and the Coast: Building Resilient Communities [Glavovic, B.,
R.C. Kay, M. Kelly, and A. Travers (eds.)]. CRC Press, London, UK (in press),
www.crcpress.com/product/isbn/9780415464871.
KCDC, 2012: Proposed Kapiti Coast District Plan, Volume 1. Notified 29 November
2012, Kapiti Coast District Council (KCDC), Paraparaumu, New Zealand, 1036 pp.
Kearney, M.R., W.P. Porter, C. Williams, S. Ritchie, and A.A. Hoffmann, 2009: Integrating
biophysical models and evolutionary theory to predict climatic impacts on
species’ ranges: the dengue mosquito Aedes aegypti in Australia. Functional
Ecology, 23(3), 528-538.
Kearney, M.R., N.J. Briscoe, D.J. Karoly, W.P. Porter, M. Norgate, and P. Sunnucks,
2010a: Early emergence in a butterfly causally linked to anthropogenic warming.
Biology Letters, 6(5), 674-677.
Kearney, M.R., B.A. Wintle, and W.P. Porter, 2010b: Correlative and mechanistic models
of species distribution provide congruent forecasts under climate change.
Conservation Letters, 3(3), 203-213.
Keith, D.A., H.R. Akcakaya, W. Thuiller, G.F. Midgley, R.G. Pearson, S.J. Phillips, H.M.
Regan, M.B. Araujo, and T.G. Rebelo, 2008: Predicting extinction risks under
climate change: coupling stochastic population models with dynamic
bioclimatic habitat models. Biology Letters, 4(5), 560-563.
Keith, D.A., S. Rodoreda, and M. Bedward, 2010: Decadal change in wetland-
woodland boundaries during the late 20
th
century reflects climatic trends.
Global Change Biology, 16(8), 2300-2306.
Kenderdine, S., 2010: Examining climate change: an Environment Court perspective.
Resource Management Theory & Practice, 6, 35-92.
Kennedy, D., L. Stocker, and G. Burke, 2010: Australian local government action on
climate change adaptation: some critical reflections to assist decision-making.
Local Environment, 15(9/10), 805-816.
Kenny, G., 2011: Adaptation in agriculture: lessons for resilience from eastern regions
of New Zealand. Climatic Change, 106(3), 441-462.
Kerr, S. and W. Zhang, 2009: Allocation of New Zealand Units within Agriculture in
the New Zealand Emissions Trading System. Motu Working Paper 09-16, Motu
Economic and Public Policy Research, Wellington, New Zealand, 68 pp.
Khalaj, B., G. Lloyd, V. Sheppeard, and K. Dear, 2010: The health impacts of heat
waves in five regions of New South Wales, Australia: a case-only analysis.
International Archives of Occupational and Environmental Health, 83(7), 833-
842.
Khan, S., 2012: Vulnerability assessments and their planning implications: a case
study of the Hutt Valley, New Zealand. Natural Hazards, 64(2), 1587-1607.
Kharin, V.V., F.W. Zwiers, X. Zhang, and M. Wehner, 2013: Changes in temperature
and precipitation extremes in the CMIP5 ensemble. Climatic Change, 119(2),
345-357.
Kiem, A.S. and E.K. Austin, 2012: Limits and Barriers to Climate Change Adaptation
for Small Inland Communities Affected by Drought. Final report to the National
Climate Change Adaptation Research Facility (NCCARF), Griffith University,
Gold Coast Campus, Southport, QLD, Australia, 37 pp.
Kiem, A.S. and E.K. Austin, 2013: Drought and the future of rural communities:
opportunities and challenges for climate change adaptation in regional Victoria,
Australia. Global Environmental Change, 23(5), 1307-1316.
Kiem, A.S., S.W. Franks, and G. Kuczera, 2003: Multi-decadal variability of flood risk.
Geophysical Research Letters, 30(2), L015992, doi:10.1029/2002GL015992.
Kiem, A.S., D.C. Verdon-Kidd, S. Boulter, and J. Palutikof, 2010: Learning from
Experience: Historical Case Studies and Climate Change Adaptation, Case Studies
of Extreme Events. Synthesis Report by the National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 33 pp.
King, D. and G. Penny, 2006: The 2
nd
Māori Climate Forum Hongoeka Marae,
Plimmerton (24 May 2006): Summary Report. NIWA Client Report: AKL-2006-
099, National Institute of Water & Atmospheric Research (NIWA) Ltd., Auckland,
New Zealand, 23 pp.
King, D., W. Tawhai, A. Skipper, and W. Iti, 2005: Anticipating Local Weather and Climate
Outcomes using Māori Environmental Indicators. NIWA Client Report: AKL-
2005-129, Prepared for FRST Programme: Adaptation to Climate Variability and
Change Human Dimensions, National Institute of Water and Atmospheric
Research (NIWA) Ltd., Auckland, New Zealand, 15 pp.
King, D., J. Goff, and A. Skipper, 2007: Māori environmental knowledge and natural
hazards in Aotearoa-New Zealand. Journal of the Royal Society of New Zealand,
37(2), 59-73.
King, D., W. Iti, and D. Hosking, 2008: Ground-Truthing Pre-Event Recovery Planning
Issues with Ngāti Rongomai. NIWA Client Report: AKL-2008-087, Prepared for
GNS Science and Ngāti Rongomai, National Institute of Water and Atmospheric
Research (NIWA) Ltd., Auckland, New Zealand, 72 pp.
King, D., G. Penny, and C. Severne, 2010: The climate change matrix facing Maori
society. In: Climate Change Adaptation in New Zealand: Future Scenarios and
Some Sectoral Perspectives [Nottage, R., D. Wratt, J. Bornman, and K. Jones
(eds.)]. New Zealand Climate Change Centre, Wellington, New Zealand, pp.
100-111.
King, D., W. Dalton, M. Home, M. Duncan, M.S. Srinivasan, J. Bind, C. Zammit, A.
McKerchar, D. Ashford-Hosking, and A. Skipper, 2012: Māori Community
Adaptation to Climate Variability and Change: Examining Risks, Vulnerability
and Adaptive Strategies with Ngāti Huirapa at Arowhenua Pā, Te-umu-kaha
(Temuka), New Zealand. NIWA Client Report: AKL2011-015, Prepared for Te
Rūnanga o Arowhenua Society Incorporated and the New Zealand Climate
Change Research Institute, Victoria University by National Institute of
Water and Atmospheric Research (NIWA) Ltd., NIWA, Auckland, New Zealand,
133 pp.
King, K.J., R.M. de Ligt, and G.J. Cary, 2011: Fire and carbon dynamics under climate
change in south-eastern Australia: insights from FullCAM and FIRESCAPE
modelling. International Journal of Wildland Fire, 20(4), 563-577.
Kingsford, R.T. (ed.), 2011: Conservation management of rivers and wetlands under
climate change. Marine & Freshwater Research, 62(3 SI), 217-327.
Kingsford, R.T. and J.E.M. Watson, 2011: Climate change in Oceania: a synthesis of
biodiversity impacts and adaptations. Pacific Conservation Biology, 17, 270-
284.
Kingsford, R.T., J.E.M. Watson, C.J. Lundquist, O. Venter, L. Hughes, E.L. Johnston, J.
Atherton, M. Gawel, D.A. Keith, B.G. Mackey, C. Morley, H.P. Possingham, B.
Raynor, H.F. Recher, and K.A. Wilson, 2009: Major conservation policy issues
for biodiversity in Oceania. Conservation Biology, 23(4), 834-840.
Kinney, P., 2012: Winter mortality in a changing climate: will it go down? Bulletin
épidémiologique hebdomadaire, 12-13, 148-151.
Kirby, M., J. Connor, R. Bark, E. Qureshi, and S. Keyworth, 2012: The Economic Impact
of Water Reductions during the Millennium Drought in the Murray-Darling
Basin. Contributed paper prepared for presentation at the 56th AARES Annual
Conference, 7-10 February 2012, Fremantle, Western Australia, Australian
Agricultural & Resource Economics Society (AARES), Canberra, ACT, Australia,
26 pp., ageconsearch.umn.edu/bitstream/124490/2/2012AC%20Kirby%20CP%
20of%20Poster2.pdf.
Kirby, M., F. Chiew, M. Mainuddin, B. Young, G. Podger, and A. Close, 2013: Drought
and climate change in the Murray-Darling Basin: a hydrological perspective.
In: Drought in Arid and Semi-Arid Environments: A Multi-Disciplinary and Cross-
Country Perspective [Schwabe, K., J. Albiac, J. Connor, R. Hassan, and L. Meza-
Gonzales (eds.)]. Springer, Dordrecht, Netherlands, pp. 281-299.
Kirono, D.G.C. and D. Kent, 2010: Assessment of rainfall and potential evaporation
from global climate models and its implications for Australian regional drought
projection. International Journal of Climatology, 31(9), 1295-1308.
1426
Chapter 25 Australasia
25
Kirono, D.G.C., D.M. Kent, K.J. Hennessy, and F.S. Mpelasoka, 2011: Characteristics
of Australian droughts under enhanced greenhouse conditions: results from 14
global climate models. Journal of Arid Environments, 75(6), 566-575.
Kirschbaum, M.U.F. and M.S. Watt, 2011: Use of a process-based model to describe
spatial variation in Pinus radiata productivity in New Zealand. Forest Ecology
and Management, 262, 1008-1019.
Kirschbaum, M.U.F., N.W.H. Mason, M.S. Watt, A. Tait, A.E. Ausseil, D.J. Palmer, and
F.E. Carswell, 2011a: Productivity Surfaces for Pinus radiata and a Range of
Indigenous Forest Species under Current Climatic Conditions. MAF Technical
Paper No: 2011/45, Full report prepared for the Ministry of Agriculture and
Forestry (MAF) by Landcare Research (CO9X0902), MAF, Wellington, New
Zealand, 119 pp.
Kirschbaum, M.U.F., D. Whitehead, S.M. Dean, P.N. Beets, J.D. Shepherd, and A.G.E.
Ausseil, 2011b: Implications of albedo changes following afforestation on the
benefits of forests as carbon sinks. Biogeosciences, 8(12), 3687-3696.
Kirschbaum, M.U.F., M.S. Watt, A. Tait, and A.-G.E. Ausseil, 2012: Future productivity
of Pinus radiata in New Zealand under expected climatic changes. Global
Change Biology, 18(4), 1342-1356.
Klamt, M., R. Thompson, and J. Davis, 2011: Early response of the platypus to climate
warming. Global Change Biology, 17(10), 3011-3018.
Kouvelis, B., C. Scott, E. Rudkin, D. Cameron, M. Harkness, A. Renata, and G. Williams,
2010: Impacts of Climate Change on Rural Water Infrastructure. Report to the
Ministry of Agriculture and Forestry (MAF) by MWH New Zealand, MAF,
Wellington, New Zealand, 134 pp.
Kuczera, G., 1987: Prediction of water yield reductions following a bushfire in ash-
mixed species eucalypt forest. Journal of Hydrology, 94(3-4), 215-236.
Kulendran, N. and L. Dwyer, 2010: Seasonal Variation versus Climate Variation for
Australian Tourism. Sustainable Tourism Cooperative Research Centre (STCRC),
Griffith University, Gold Coast Campus, Southport, QLD, Australia, 27 pp.
Kuleshov, Y., R. Fawcett, L. Qi, B. Trewin, D. Jones, J. McBride, and H. Ramsay, 2010:
Trends in tropical cyclones in the South Indian Ocean and the South Pacific
Ocean. Journal of Geophysical Research: Atmospheres, 115, D01101,
doi:10.1029/2009JD012372.
Kuruppu, N., J. Murta, P. Mukheibir, J. Chong, and T. Brennan, 2013: Understanding
the Adaptive Capacity of Australian Small-To-Medium Enterprises to Climate
Change and Variability: Final Report. National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 129 pp.
Kwan, D., H. Marsh, and S. Delean, 2006: Factors influencing the sustainability of
customary dugong hunting by a remote indigenous community. Environmental
Conservation, 33(2), 164-171.
Lal, A., S. Hales, N. French, M.G. Baker, 2012: Seasonality in human zoonotic enteric
diseases: a systematic review. PLoS ONE, 7(4), e31883, doi:10.1371/journal.
pone.0031883.
Langton, M., M. Parsons, S. Leonard, K. Auty, D. Bell, C.P. Burgess, S. Edwards, R.
Howitt, S. Jackson, V. McGrath, and J. Morrison, 2012: National Climate Change
Adaptation Research Plan: Indigenous Communities. NCCARF Publication
11/12, National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 48 pp.
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.
Laurance, W.F., 2008: Global warming and amphibian extinctions in eastern Australia.
Austral Ecology, 33(1), 1-9.
Laurance, W.F., B. Dell, S.M. Turton, M.J. Lawes, L.B. Hutley, H. McCallum, P. Dale, M.
Bird, G. Hardy, G. Prideaux, B. Gawne, C.R. McMahon, R. Yu, J.M. Hero, L.
Schwarzkop, A. Krockenberger, M. Douglas, E. Silvester, M. Mahony, K. Vella, U.
Saikia, C.H. Wahren, Z.H. Xu, B. Smith, and C. Cocklin, 2011: The 10 Australian
ecosystems most vulnerable to tipping points. Biological Conservation, 144(5),
1472-1480.
Lawrence, A., T. Baker, and C.E. Lovelock, 2012: Optimising and Managing Coastal
Carbon: Comparative Sequestration and Mitigation Opportunities across
Australia’s Landscapes and Land Uses: Final Report. FRDC Report 2011/084,
Prepared for the Fisheries Research and Development Corporation (FRDC) by
TierraMar Consulting Pty Ltd., FRDC, Deakin, ACT, Australia, 66 pp.
Lawrence, J. and S. Allan, 2009: A Strategic Framework and Practical Options for
Integrating Flood Risk Management – To Reduce Flood Risk and the Effects of
Climate Change. Prepared for the Ministry for the Environment (MfE) by PS
Consulting Ltd. and MWH New Zealand (MWH), MfE, Wellington, New Zealand,
79 pp.
Lawrence, J., A. Reisinger, B. Mullan, and B. Jackson, 2013a: Exploring climate change
uncertainties to support adaptive management of changing flood-risk.
Environmental Science & Policy, 33, 133-142.
Lawrence, J., F. Sullivan, A. Lash, G. Ide, C. Cameron, and L. McGlinchey, 2013b: Adapting
to changing climate risk by local government in New Zealand: institutional
practice barriers and enablers. Local Environment: The International Journal of
Justice and Sustainability, doi:10.1080/13549839.2013.839643.
Leblanc, M., S. Tweed, A. Van Dijk, and B. Timbal, 2012: A review of historical and
future hydrological changes in the Murray-Darling Basin. Global and Planteray
Change, 80-81, 226-246.
Leitch, A.M. and C.J. Robinson, 2012: Shifting sands: uncertainty and a local
community response to sea level rise policy in Australia. In: Risk and Social Theory
in Environmental Management: From Uncertainty to Local Action [Measham,
T. and S. Lockie (eds.)]. Commonwealth Scientific and Industrial Research
Organization (CSIRO) Publishing, Collingwood, VIC, Australia, pp. 117-131.
Lennox, J., W. Proctor, and S. Russell, 2011: Structuring stakeholder participation in
New Zealand’s water resource governance. Ecological Economics, 70(7), 1381-
1394.
Leonard, M., S. Westra, A. Phatak, M. Lambert, B. van den Hurk, K. McInnes, J. Risbey,
S. Schuster, D. Jakob, and M. Stafford-Smith, 2013: A compound event framework
for understanding extreme impacts. Wiley Interdisciplinary Reviews: Climate
Change, 5(1), 113-128.
Leonard, S., J. Mackenzie, F. Kofod, M. Parsons, M. Langton, P. Russ, L. Ormond-Parker,
K. Smith, and M. Smith, 2013: Indigenous Climate Change Adaptation in the
Kimberley Region of North-Western Australia: Learning from the Past, Adapting
in the Future: Identifying Pathways to Successful Adaptation in Indigenous
Communities. National Climate Change Adaptation Research Facility (NCCARF),
Griffith University, Gold Coast Campus, Southport, QLD, Australia, 131 pp.
Leslie, L., M. Leplastrier, and B. Buckley, 2008: Estimating future trends in severe
hailstorms over the Sydney Basin: a climate modelling study. Atmospheric
Research, 87(1), 37-51.
Lett, R., R. Morden, C. McKay, T. Sheedy, M. Burns, and D. Brown, 2009: Farm dam
interception in the Campaspe Basin under climate change. In: H
2
009: 32
nd
Hydrology and Water Resources Symposium, 30 November-3 December 2009,
Newcastle, Australia. Engineers Australia, Barton, ACT, Australia, pp 1194-1204.
Leviston, Z., A. Leitch, M. Greenhill, R. Leonard, and I. Walker, 2011: Australians
Views of Climate Change. Commissioned by the Garnaut Climate Change
Review, Commonwealth Scientific and Industrial Research Organization
(CSIRO), Canberra, ACT, Australia, 21 pp.
Leviston, Z., I. Walker, and S. Morwinski, 2012: Your opinion on climate change might
not be as common as you think. Nature Climate Change, 3, 334-337.
Lewandowsky, S., 2011: Popular consensus: climate change is set to continue.
Psychological Science, 22(4), 460-463.
LGNZ, 2008: Local Government New Zealand’s Submission to the Board of Inquiry
on the “Proposed Coastal Policy Statement”. Submission to the Department
of Conservation by Local Government New Zealand (LGNZ), Wellington, New
Zealand, 38 pp.
Lieffering, M., P.C.D. Newton, F.Y. Li, and R. Vibart, 2012: Hill country sheep and beef:
impacts and adaptation to climate changeIn: Enhanced Climate Change Impact
and Adaptation Evaluation: A Comprehensive Analysis of New Zealand’s Land-
Based Primary Sectors [Clark, A. and R. Nottage (eds.)]. Ministry for Primary
Industries, Wellington, New Zealand, pp. 145-188.
Lindenmayer, D.B., 2007: On Borrowed Time: Australia’s Environmental Crisis and
What We Must Do About It. Penguin Books in association with Commonwealth
Scientific and Industrial Research Organization (CSIRO) Publishing, Camberwell,
VIC, Australia,138 pp.
Ling, N., 2010: Socio-economic drivers of freshwater fish declines in a changing
climate: a New Zealand perspective. Journal of Fish Biology, 77(8), 1983-1992.
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, K. Ridgway, A.J. Hobday, and M. Haddon, 2009: Climate-
driven range extension of a sea urchin: inferring future trends by analysis of
recent population dynamics. Global Change Biology, 15(3), 719-731.
1427
Australasia Chapter 25
25
Linnenluecke, M.K., A. Stathakis, and A. Griffiths, 2011: Firm relocation as adaptive
response to climate change and weather extremes. Global Environmental
Change, 21(1), 123-133.
Loechel, B., J.H. Hodgkinson, K. Moffat, S. Crimp, A. Littleboy, and M. Howden, 2010:
Goldfields-Esperance Regional Mining Climate Vulnerability Workshop: Report
on Workshop Outcomes. Commonwealth Scientific and Industrial Research
Organization (CSIRO), Canberra, ACT, Australia, 41 pp.
Loechel, B., J. Hodgkinson, and K. Moffat, 2013: Climate change adaptation in
Australian mining communities: comparing mining company and local
government views and activities. Climatic Change, 119(2), 465-477.
Lough, J.M., 2008: Shifting climate zones for Australia’s tropical marine ecosystems.
Geophysical Research Letters, 35(14), L14708, doi:10.1029/2008GL034634.
Lough, J.M. and A.J. Hobday, 2011: Observed climate change in Australian marine
and freshwater environments. Marine and Freshwater Research, 62(9), 984-
999.
Loughnan, M.E., N. Nicholls, and N.J. Tapper, 2010: The effects of summer temperature,
age and socioeconomic circumstance on acute myocardial infarction admissions
in Melbourne, Australia. International Journal of Health Geographics, 9, 41,
doi:10.1186/1476-072X-9-41.
Lovelock, C.E., G. Skilleter, and N. Saintilan, 2009: Tidal wetlands and climate change.
In: A Marine Climate Change Impacts and Adaptation Report Card for Australia
2009 [Poloczanska, E.S., A.J. Hobday, and A.J. Richardson (eds.)]. NCCARF
Publication 05/09, National Climate Change Adaptation Research Facility
(NCCARF), Griffith University, Gold Coast Campus, Southport, QLD, Australia,
32 pp.
Low Choy, D., S. Serrao-Neumann, F. Crick, G. Schuch, M. Sanò, R.v. Staden, O. Sahin,
B. Harman, and S. Baum, 2012: Adaptation Options for Human Settlements in
South East Queensland – Main Report. Unpublished Report for the South East
Queensland Climate Adaptation Research Initiative (SEQ CARI), Griffith
University, Nathan, QLD, Australia, 210 pp.
Lucas, C., K.J. Hennessy, and J.M. Bathols, 2007: Bushfire Weather in Southeast
Australia: Recent Trends and Projected Climate Change Impacts. Consultancy
Report prepared for The Climate Institute of Australia by the Bushfire Cooperative
Research Centre (Bushfire CRC) primarily from standard observations made by
the Australian Bureau of Meteorology with southeastern Australia climate
change projections from CSIRO, Bushfire CRC, Melbourne, VIC, Australia, 80
pp.
Lundquist, C.J., D. Ramsay, R. Bell, A. Swales, and S. Kerr, 2011: Predicted impacts of
climate change on New Zealand’s biodiversity. Pacific Conservation Biology,
17(3), 179-191.
Lunt, I.D., M. Byrne, J.J. Hellmann, N.J. Mitchell, S.T. Garnett, M.W. Hayward, T.G.
Martin, E. McDonald-Madden, S.E. Williams, and K.K. Zander, 2013: Using
assisted colonisation to conserve biodiversity and restore ecosystem function
under climate change. Biological Conservation, 157, 172-177.
Luo, Q., W. Bellotti, M. Williams, and E. Wang, 2009: Adaptation to climate change of
wheat growing in South Australia: analysis of management and breeding
strategies. Agriculture, Ecosystems & Environment, 129(1-3), 261-267.
LVRC, 2012: Lockyer Valley Relocation Policy. Lockyer Valley Regional Council (LVRC),
Gatton, QLD, Australia,13 pp.
LWF, 2010: Report of the Land and Water Forum: A Fresh Start for Freshwater. Land
and Water Forum (LWF), Wellington, New Zealand, 68 pp.
Mac Nally, R., G. Horrocks, H. Lada, P.S. Lake, J.R. Thomson, and A.C. Taylor, 2009:
Distribution of anuran amphibians in massively altered landscapes in south-
eastern Australia: effects of climate change in an aridifying region. Global
Ecology and Biogeography, 18(5), 575-585.
Macintosh, A., 2013: Coastal climate hazards and urban planning: how planning
responses can lead to maladaptation. Mitigation and Adaptation Strategies for
Global Change, 18(7), 1035-1055.
Macintosh, A., A. Foerster, and J. McDonald, 2013: Limp, Leap or Learn? Developing
Legal Frameworks for Climate Change Adaptation Planning in Australia: Final
Report. National Climate Change Adaptation Research Facility (NCCARF),
Griffith University, Gold Coast Campus, Southport, QLD, Australia, 267 pp.
Mackey, B.G., J.E.M. Watson, G. Hope, and S. Gilmore, 2008: Climate change,
biodiversity conservation, and the role of protected areas: an Australian
perspective. Biodiversity, 9, 11-18.
Madin, E.M.P., N.C. Ban, Z.A. Doubleday, T.H. Holmes, G.T. Pecl, and F. Smith, 2012:
Socio-economic and management implications of range-shifting species in
marine systems. Global Environmental Change: Human and Policy Dimensions,
22(1), 137-146.
Maggini, R., H. Kujala, M.F.J. Taylor, J.R. Lee, H.P. Possingham, B.A. Wintle, and R.A.
Fuller, 2013: Protecting and Restoring Habitat to Help Australia’s Threatened
Species Adapt to Climate Change: Final Report. National Climate Change
Adaptation Research Facility (NCCARF), Griffith University, Gold Coast Campus,
Southport, QLD, Australia, 53 pp.
Maloney, S.K. and C.F. Forbes, 2011: What effect will a few degrees of climate change
have on human heat balance? Implications for human activity. International
Journal of Biometeorology, 55(2), 147-160.
Mankad, A. and S. Tapsuwan, 2011: Review of socio-economic drivers of community
acceptance and adoption of decentralised water systems. Journal of Environmental
Management, 92(3), 380-391.
Mapstone, B.D., P. Appleford, K. Broderick, R. Connolly, J. Higgins, A. Hobday, T.P.
Hughes, P.A. Marshall, J. McDonald, and M. Waschka, 2010: National Climate
Change Adaptation Research Plan for Marine Biodiversity and Resources.
National Climate Change Adaptation Research Facility, Griffith University, Gold
Coast Campus, Southport, QLD, Australia, 68 pp.
Marcar, N.E., R.G. Benyon, P.J. Polglase, K.I. Paul, S. Theiveyanathan, and L. Zhang,
2006: Predicting the Hydrological Impacts of Bushfire and Climate Change in
Forested Catchments of the River Murray Uplands: A Review. Commonwealth
Scientific and Industrial Research Organization (CSIRO) Water for a Healthy
Country National Research Flagship, Canberra, ACT, Australia, 38 pp.
Marshall, N.A., 2010: Understanding social resilience to climate variability in primary
enterprises and industries. Global Environmental Change, 20(1), 36-43.
Marshall, P.A. and J.E. Johnson, 2007: Chapter 24: The Great Barrier Reef and climate
change: vulnerability and management implications. 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) and Australian
Greenhouse Office, Townsville, QLD, Australia, pp. 773-801.
Maru, Y.T., J. Langridge, and B.B. Lin, 2011: Current and Potential Applications of
Typologies in Vulnerability Assessments and Adaptation Science. CSIRO Climate
Adaptation Flagship Working Paper No. 7, Commonwealth Scientific and
Industrial Research Organization (CSIRO), Melbourne, VIC, Australia, 25 pp.
Maru, Y.T., V. Chewings, and A. Sparrow, 2012: Climate Change Adaptation, Energy
Futures and Carbon Economies in Remote Australia: A Review of the Current
Literature, Research and Policy. CRC-REP Working Paper CW005, Ninti One Ltd.,
Alice Springs, NT, Australia, 84 pp.
Mason, M. and K. Haynes, 2010: Adaptation Lessons from Cyclone Tracy. National
Climate Change Adaptation Research Facility (NCCARF), Griffith University,
Gold Coast Campus, Southport, QLD, Australia, 82 pp.
Mason, M., K. Haynes, and G. Walker, 2013: Cyclone Tracy and the road to improving
wind resistant design. In: Natural Disasters and Adaptation to Climate Change
[Boulter, S., J. Palutikof, D. Karoly, and D. Guitart (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 137-149.
Matusik, G., K.X. Ruthrof, N.C. Brouwers, B. Dell, and G.S.J. Hardy, 2013: Sudden
forest canopy collapse corresponding with extreme drought and heat in a
Mediterranean-type euclaypt forest in southwestern Australia. European Journal
of Forest Research, 132, 497-510.
Maunsell and CSIRO, 2008: Impacts of Climate Change on Infrastructure in Australia
and CGE Model Inputs. Report prepared by Maunsell Australia Pty Ltd in
association with Commonwealth Scientific and Industrial Research Organization
(CSIRO) Sustainable Ecosystems for the Garnaut Climate Change Secretariat,
Garnaut Climate Change Review, Canberra, ACT, Australia, 130 pp.
Mayor, K. and R.S.J. Tol, 2007: The impact of the UK aviation tax on carbon dioxide
emissions and visitor numbers. Transport Policy, 14(6), 507-513.
McAdam, J., 2010: Disappearing states’, statelessness and the boundaries of
international law. In: Climate Change and Displacement [McAdam, J. (ed.)].
Hart Publishing, Oxford, UK, pp. 105-130.
McAlpine, C.A., J. Syktus, R.C. Deo, P.J. Lawrence, H.A. McGowan, I.G. Watterson,
and S.R. Phinn, 2007: Modeling the impact of historical land cover change on
Australias regional climate. Geophysical Research Letters, 34(22), L22711,
doi:10.1029/2007GL031524.
McAlpine, C.A., J. Syktus, J.G. Ryan, R.C. Deo, G.M. McKeon, H.A. McGowan, and S.R.
Phinn, 2009: A continent under stress: interactions, feedbacks and risks
associated with impact of modified land cover on Australia’s climate. Global
Change Biology, 15(9), 2206-2223.
McBride, J., 2012: The Estimated Cost of Tropical Cyclone Impacts in Western
Australia. Technical Report for the Indian Ocean Climate Initiative (IOCI) Stage
3. Project 2.2: Tropical Cyclones in the North West, Indian Ocean Climate
Initiative (Government of Western Australia, Commonwealth Scientific and
1428
Chapter 25 Australasia
25
Industrial Research Organization (CSIRO), Bureau of Meteorology (BoM)), Perth,
WA, Australia, 53 pp.
McBride, J. and N. Nicholls, 1983: Seasonal relationships between Australian rainfall
and the Southern Oscillation. Monthly Weather Review, 111(10), 1998-2004.
McCallum, J.L., R.S. Crosbie, G.R. Walker, and W.R. Dawes, 2010: Impacts of climate
change on groundwater in Australia: a sensitivity analysis of recharge.
Hydrogeology Journal, 18(7), 1625-1638.
McCleave, J., S. Espiner, and K. Booth, 2006: The New Zealand people-park relationship:
an exploratory model. Society and Natural Resources, 19(6), 547-561.
McDonald, J., 2010: Paying the price of climate change adaptation: compensation
for climate change impacts. In: Adaptation to Climate Change: Law and Policy
[Bonyhady, T., A. Macintosh, and J. McDonald (eds.)]. Federation Press, Annandale,
NSW, Australia, pp. 234-264.
McDonald, J., 2011: The role of law in adapting to climate change. Wiley
Interdisciplinary Reviews: Climate Change, 2, 283-295.
McDonald, J., 2013: Creating legislative frameworks for adaptation. In: Climate
Adaptation Futures [Palutikof, J., S.L. Boulter, A.J. Ash, M. Stafford-Smith, M.
Parry, M. Waschka, and D. Guitart (eds.)]. John Wiley & Sons, Brisbane, QLD,
Australia, pp. 126-132.
McGlone, M.S. and S. Walker, 2011: Potential Effects of Climate Change on New
Zealand’s Terrestrial Biodiversity and Policy Recommendations for Mitigation,
Adaptation and Research. Science for Conservation Monograph Series No. 312,
Department of Conservation (DOC), Wellington, New Zealand, 77 pp.
McGlone, M., S. Walker, R. Hay, and J. Christie, 2010: Climate change, natural systems
and their conservation in New Zealand. In: Climate Change Adaptation in New
Zealand: Future Scenarios and Some Sectoral Perspectives [Nottage, R., D.
Wratt, J. Bornman, and K. Jones (eds.)]. New Zealand Climate Change Centre,
c/o National Institute of Water and Atmospheric Research (NIWA), Wellington,
New Zealand, pp. 82-99.
McHenry, M.P., 2009: Agricultural bio-char production, renewable energy generation
and farm carbon sequestration in Western Australia: certainty, uncertainty and
risk. Agriculture, Ecosystems & Environment, 129(1-3), 1-7.
McInnes, K.L., D.J. Abbs, and J.A. Bathols, 2005: Climate Change in Eastern Victoria.
Stage 1 Report: The Effect of Climate Change on Coastal Wind and Weather
Patterns. Consultancy Report for the Gippsland Coastal Board by the Common-
wealth Scientific and Industrial Research Organization (CSIRO) Marine and
Atmospheric Research, Aspendale, VIC, Australia, 26 pp.
McInnes, K.L., I. Macadam, G.D. Hubbert, and J.G. O’Grady, 2009: A modelling
approach for estimating the frequency of sea level extremes and the impact of
climate change in southeast Australia. Natural Hazards, 51(1), 115-137.
McInnes, K.L., T.A. Erwin, and J.M. Bathols, 2011a: Global Climate Model projected
changes in 10 m wind speed and direction due to anthropogenic climate
change. Atmospheric Science Letters, 12(4), 325-333.
McInnes, K.L., I. Macadam, G.D. Hubbert, and J.G. O’Grady, 2011b: An assessment of
current and future vulnerability to coastal inundation due to sea level extremes in
Victoria, southeast Australia. International Journal of Climatology, 33(1), 33-47.
McInnes, K.L., J.G. O’Grady, M. Hemer, I. Macadam, D.J. Abbs, C.J. White, S.P. Corney,
M.R. Grose, G.K. Holz, S.M. Gaynor, and N.L. Bindoff, 2012: Climate Futures for
Tasmania: Extreme Tide and Sea-Level Events Technical Report. Antarctic Climate
and Ecosystems Cooperative Research Centre (CRC), Hobart, TAS, Australia, 42 pp.
McIntyre, S., 2011: Ecological and anthropomorphic factors permitting low-risk
assisted colonization in temperate grassy woodlands. Biological Conservation,
144(6), 1781-1789.
McIntyre-Tamwoy, S. and A. Buhrich, 2011: Lost in the wash: predicting the impact
of losing Aboriginal coastal sites in Australia. International Journal of Climate
Change: Impacts and Responses, 3(1), 53-66.
McKechnie, A.E. and B.O. Wolf, 2010: Climate change increases the likelihood of
catastrophic avian mortality events during extreme heat waves. Biology Letters,
6(2), 253-256.
McKeon, G.M., G.S. Stone, J.I. Syktus, J.O. Carter, N.R. Floof, D.G. Ahrens, D.N. Bruget,
C.R. Chilcott, D.H. Cobon, R.A. Cowley, S.J. Crimp, G.W. Fraser, S.M. Howden,
P.W. Johnston, J.G. Ryan, C.J. Stokes, and K.A. Day, 2009: Climate Change
Impacts on Australia’s Rangeland Livestock Carrying Capacity: A Review of
Challenges. Report for Land & Water Australia Senior Research Fellowship
(QNR46), Land & Water Australia, Canberra, ACT, Australia, 69 pp.
McKerchar, A. and B. Mullan, 2004: Seasonal Inflow Distributions for New Zealand
Hydroelectric Power Stations. NIWA Client Report: CHC2004-131, Prepared for
the Ministry of Economic Development (MED) by the National Institute of Water
and Atmospheric Research (NIWA), MED, Wellington, New Zealand, 12 pp.
McKerchar, A.I., J.A. Renwick, and J. Schmidt, 2010: Diminishing streamflows on the
east coast of the South Island New Zealand and linkage to climate variability
and change. Journal of Hydrology (NZ), 49, 1-14.
Mcleod, D.J., G.M. Hallegraeff, G.W. Hosie, and A.J. Richardson, 2012: Climate-driven
range expansion of the red-tide dinoflagellate Noctiluca scintillans into the
Southern Ocean. Journal of Plankton Research, 34(4), 332-337.
McManus, P., J. Walmsley, N. Argent, S. Baumc, L. Bourked, J. Martine, B. Pritchard,
and T. Sorensen, 2012: Rural community and rural resilience: what is important
to farmers in keeping their country towns alive? Journal of Rural Studies, 28(1),
20-29.
McMichael, A., H.J. Weaver, H. Berry, P. Beggs, B.J. Currie, J. Higgins, B. Kelly, J. McDonald,
and S. Tong, 2009: National Climate Change Adaptation Plan for Human Health.
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 64 pp.
McMillan, H., B. Jackson, and S. Poyck, 2010: Flood Risk Under Climate Change:A
Framework for Assessing the Impacts of Climate Change on River Flow and
Floods, using Dynamically Down-Scaled Climate Scenarios. A case-study for the
Uawa (East Cape) and Waihou (Northland) catchments, NIWA Report CHC2010-
033, Prepared for the Ministry for Agriculture and Forestry (MAF) by the
National Institute of Water and Atmospheric Research (NIWA), MAF, Wellington,
New Zealand, 53 pp.
McMillan, H., M. Duncan, G. Smart, J. Sturman, S. Poyck, S. Reese, A. Tait, E. Hreinsson,
and J. Walsh, 2012: The urban impacts toolbox: an example of modelling the
effect of climate change and sea level rise on future flooding. Weather and
Climate, 32(2), 21-39.
McNamara, K.E., S.G. Smithers, R. Westoby, and K. Parnell, 2011: Limits to Climate
Change Adaptation for Low-Lying Communities in the Torres Strait: Final Report.
Prepared for the National Climate Change Adaptation Research Facility
(NCCARF) by James Cook University, NCCARF, Griffith University, Gold Coast
Campus, Southport, QLD, Australia, 87 pp.
McVicar, T.R., T.G. Van Niel, L.T. Li, M.L. Roderick, D.P. Rayner, L. Ricciardulli, and R.J.
Donohue, 2008: Wind speed climatology and trends for Australia, 1975-2006:
capturing the stilling phenomenon and comparison with near-surface reanalysis
output. Geophysical Research Letters, 35(20), L035627, doi:10.1029/
2008GL035627.
McVicar, T.R., R.J. Donohue, A.P. O’Grady, and L.T. Li, 2010: The Effects of Climatic
Changes on Plant Physiological and Catchment Ecohydrological Processes in
the High-Rainfall Catchments of the Murray-Darling Basin: A Scoping Study.
Prepared for the Murray-Darling Basin Authority (MDBA) by the Commonwealth
Scientific and Industrial Research Organization (CSIRO) Water for a Healthy
Country National Research Flagship, MDBA, Canberra, ACT, Australia, 95
pp.
MDBA, 2011: Proposed Basin Plan, Water Act 2007. MDBA Publication No. 192/11,
Draft plan prepared for the purpose of consultation for the Commonwealth of
Australia by the Murray-Darling Basin Authority (MDBA), MDBA, Canberra, ACT,
Australia, 210 pp.
MDBA, 2012: Proposed Basin Plan A Revised Draft. MDBA Publication No. 57/12,
Prepared by Murray-Darling Basin Authority (MDBA), as required by subsection
43A(2) of the Water Act 2007, MDBA, Canberra, ACT, Australia, 215 pp.
MDBC, 2007: Impact of the 2003 Alpine Bushfires on Streamflow: Broadscale Water
Yield Assessment. MDBC Publication No. 21/08, Prepared by Sinclair Knight
Merz (SKM) for the Victorian Department of Sustainability and Environment
and the Murray-Darling Basin Commission (MDBC), MDBC, Canberra, ACT,
Australia, 118 pp.
Measham, T., B. Preston, T. Smith, C. Brooke, R. Gorddard, G. Withycombe, and C.
Morrison, 2011: Adapting to climate change through local municipal planning:
barriers and challenges. Mitigation and Adaptation Strategies for Global
Change, 16(8), 889-909.
MED, 2011: New Zealand’s Energy Outlook 2011: Reference Scenario and Sensitivity
Analysis. Ministry of Economic Development (MED), Wellington, New Zealand,
12 pp.
Medlyn, B.E., R.A. Duursma, and M.J.B. Zeppel, 2011a: Forest productivity under climate
change: a checklist for evaluating model studies. Wiley Interdisciplinary
Reviews: Climate Change, 2(3), 332-355.
Medlyn, B.E., M. Zeppel, N.C. Brouwers, K. Howard, E. O’Gara, G. Hardy, T. Lyons, L.
Li, and B. Evans, 2011b: Biophysical Impacts of Climate Change on Australia’s
Forests. Contribution of Work Package 2 to the Forest Vulnerability Assessment.
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 189 pp.
1429
Australasia Chapter 25
25
Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A.
Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper, I.G. Watterson, A.J. Weaver,
and Z.-C. Zhao, 2007: Global climate projections. 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 [Solomon, S., 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. 747-
845.
Melbourne Water, 2010: Melbourne Water Responding to a Changing Climate.
Melbourne Water, Melbourne, VIC, Australia, 19 pp.
Melloy, P., G. Hollaway, J. Luck, R. Norton, E. Aitken, and S. Chakraborty, 2010:
Production and fitness of Fusarium pseudograminearum inoculum at elevated
carbon dioxide in FACE. Global Change Biology, 16(12), 3363-3373.
Memon, A. and P. Skelton, 2007: Institutional arrangements and planning practices
to allocate freshwater resources in New Zealand: a way forward. New Zealand
Journal of Environmental Law, 11, 241-277.
Memon, A., B. Painter, and E. Weber, 2010: Enhancing potential for integrated
catchment management in New Zealand: a multi-scalar, strategic perspective.
Australasian Journal of Environmental Management, 17(1), 35-44.
Mendham, E. and A. Curtis, 2010: Taking over the reins: trends and impacts of changes
in rural property ownership. Society & Natural Resources, 23(7), 653-668.
Menendez, M. and P.L. Woodworth, 2010: Changes in extreme high water levels
based on a quasi-global tide-gauge data set. Journal of Geophysical Research:
Oceans, 115(C10), C10011, doi:10.1029/2009JC005997.
Meyssignac, B. and A. Cazenave, 2012: Sea level: a review of present-day and
recent-past changes and variability. Journal of Geodynamics, 58, 96-109.
MfE, 2007a: Environment New Zealand 2007. Publication No. ME 847, Ministry for
the Environment (MfE), Wellington, New Zealand, 456 pp.
MfE, 2007b: Consultation with ori on Climate Change: Hui Report. Publication
No. ME 830, Ministry for the Environment (MfE), Wellington, New Zealand, 135
pp.
MfE, 2008a: Preparing for Climate Change: A Guide for Local Government in New
Zealand. Publication No. ME 891, Ministry for the Environment, Wellington,
New Zealand, 38 pp.
MfE, 2008b: Climate Change Effects and Impacts Assessment: A Guidance Manual
for Local Government in New Zealand (2
nd
edition). Publication No. ME 870,
Prepared by Mullan, B., D. Wratt, S. Dean, M. Hollis, S. Allan, T. Williams, G. Kenny,
and MfE, Ministry for the Environment (MfE), Wellington, New Zealand, 149
pp.
MfE, 2008c: Climate Change and Long-Term Council Community Planning.
Publication No. ME 911, Ministry for the Environment (MfE), Wellington, New
Zealand, 18 pp.
MfE, 2008d: Coastal Hazards and Climate Change: A Guidance Manual for Local
Government in New Zealand (2nd edition). Publication No. ME 892, Revised
by Ramsay, D. and R. Bell, (NIWA), Prepared for Ministry for the Environment
(MfE), MfE, Wellington, New Zealand, 129 pp.
MfE, 2010a: Legally Protected Conservation Land in New Zealand: Environmental
Snapshot April 2010. Publication No. INFO 492, Ministry for the Environment
(MfE), Wellington, New Zealand, 7 pp.
MfE, 2010b: Tools for Estimating the Effects of Climate Change on Flood Flow: A
Guidance Manual for Local Government in New Zealand. Publication No. ME
1013, Prepared by Woods, R., A.B. Mullan, G. Smart, H. Rouse, M. Hollis, A.
McKerchar, R. Ibbitt, S. Dean, and D. Collins (NIWA) for the Ministry for the
Environment (MfE), MfE, Wellington, New Zealand, 63 pp.
MfE, 2011: National Policy Statement for Freshwater Management 2011. Ministry
for the Environment (MfE), Wellington, New Zealand, 11 pp.
MfE, 2013: New Zealand’s Greenhouse Gas Inventory 1990-2011. Publication No.
ME 1113. Submitted to the United Nations Framework Convention on Climate
Change (UNFCCC) by the Ministry for the Environment (MfE), MfE, Wellington,
New Zealand, 424 pp.
Milfont, T., 2012: The interplay between information level, perceived efficacy and
concern about global warming and climate change: a one-year longitudinal
study. Risk Analysis, 32(6), 1003-1020.
Milfont, T., N. Harré, C. Sibley, and J. Duckitt, 2012: The climate change dilemma:
examining the association between parental status and political party support.
Journal of Applied Social Pscyhology, 42(10), 2386-2410.
Miller, G.M., S.A. Watson, J.M. Donelson, M.I. McCormick, and P.L. Munday, 2012:
Parental environment mediates impacts of increased carbon dioxide on a coral
reef fish. Nature Climate Change, 2(12), 858-861.
Miller, K.J., A.A. Rowden, A. Williams, and V. Haussermann, 2011: Out of their depth?
Isolated deep populations of the cosmopolitan coral Desmophyllum dianthus
may be highly vulnerable to environmental change. PLoS ONE, 6(5),
doi:10.1371/journal.pone.0019004.
Mills, R., 2009: Way ahead. Rio Tinto Review, March 2009, 15-16.
Min, S.-K., W. Cai, and P. Whetton, 2013: Influence of climate variability on seasonal
extremes over Australia. Journal of Geophysical Research: Atmospheres, 118(2),
643-654.
Minister of Conservation, 2010: New Zealand Coastal Policy Statement. Department
of Conservation, Wellington, New Zealand, 28 pp.
Mitchell, N.J., F.W. Allendorf, S.N. Keall, C.H. Daugherty, and N.J. Nelson, 2010:
Demographic effects of temperature-dependent sex determination: will tuatara
survive global warming? Global Change Biology, 16(1), 60-72.
Moise, A.F. and D.A. Hudson, 2008: Probabilistic predictions of climate change for
Australia and southern Africa using the reliability ensemble average of IPCC
CMIP3 model simulations. Journal of Geophysical Research: Atmospheres,
113(D15), D009250, doi:10.1029/2007JD009250.
Moon, J.W. and S.-H. Han, 2011: Thermostat strategies impact on energy consumption
in residential buildings. Energy and Buildings, 43, 338-346.
Moore, A.D. and A. Ghahramani, 2013: Climate change and broadacre livestock
production across southern Australia. 1. Impacts of climate change on pasture
and livestock productivity, and on sustainable levels of profitability. Global
Change Biology, 19(5), 1440-1455.
Moore, J.A.Y., L.M. Bellchambers, M.R. Depczynski, R.D. Evans, S.N. Evans, S.N. Field,
K.J. Friedman, J.P. Gilmour, T.H. Holmes, R. Middlebrook, B.T. Radford, T. Ridgway,
G. Shedrawi, H. Taylor, D. Thomson, and S.K. Wilson, 2012: Unprecedented mass
bleaching and loss of coral across 12° of latitude in Western Australia in 2010-
2011. PLoS ONE, 7(12), e51807, doi:10.1371/journal.pone.0051807.
Morrison, C. and C.M. Pickering, 2011: Climate Change Adaptation in the Australian
Alps: Impacts, Strategies, Limits and Management. National Climate Change
Adaptation Research Facility (NCCARF), Griffith University, Gold Coast Campus,
Southport, QLD, Australia, 78 pp.
Morrongiello, J.R., S.J. Beatty, J.C. Bennett, D.A. Crook, D.N.E.N. Ikedife, M.J. Kennard,
A. Kerezsy, M. Lintermans, D.G. McNeil, B.J. Pusey, and T. Rayner, 2011: Climate
change and its implications for Australia’s freshwater fish. Marine and
Freshwater Research, 62, 1082-1098.
Mortimer, E., A. Bergin, and R. Carter, 2011: Sharing Risk: Financing Australia’s Disaster
Resilience. Special Report, February 2011 issue 37, Australian Strategic Policy
Institute (ASPI), Barton, ACT, Australia, 24 pp.
Mukheibir, P., N. Kuruppu, A. Gero, and J. Herriman, 2013: Cross-Scale Barriers to
Climate Change Adaptation in Local Government, Australia: Final Report.
NCCARF Publication (34/12), National Climate Change Adaptation Research
Facility (NCCARF), Griffith University, Gold Coast Campus, Southport, QLD,
Australia, 95 pp.
Mulet-Marquis, S. and J.R. Fairweather, 2008: Rural Population and Farm Labour
Change. Research Report No. 300, the Agribusiness and Economics Research
Unit (AERU), Lincoln University, Lincoln, New Zealand, 34 pp.
Mullan, B., 1995: On the linearity and stability of Southern Oscillation-climate
relationships for New Zealand. International Journal of Climatology, 15(12),
1365-1386.
Mullan, B., M.J. Salinger, C.S. Thompson, and A.S. Porteous, 2001: The New Zealand
climate: present and future. In: The Effects of Climate Change and Variation in
New Zealand: An Assessment using the Climpacts System [Warrick, R.A., G.J.
Kenny, and J.J. Harman (eds.)]. International Global Change Institute (IGCI),
University of Waikato, Hamilton, New Zealand, pp. 11-31.
Mullan, B., A. Porteous, D. Wratt, and M. Hollis, 2005: Changes in Drought Risk with
Climate Change. NIWA Client Report WLG2005-23, Prepared for the Ministry
for the Environment (MfE): NZ Climate Change Office and the Ministry of
Agriculture and Forestry by the National Institute of Water and Atmospheric
Research (NIWA), MfE, Wellington, New Zealand, 58 pp.
Mullan, B., S.J. Stuart, M.G. Hadfield, and M.J. Smith, 2010: Report on the Review of
NIWA’s ‘Seven-Station’ Temperature Series. NIWA Information Series No. 78,
2
nd
edn., December, 2010, National Institute of Water and Atmospheric Research
(NIWA), Wellington, New Zealand, 175 pp.
Mullan, B., T. Carey-Smith, G.M. Griffiths, and A. Sood, 2011: Scenarios of Storminess
and Regional Wind Extremes under Climate Change. NIWA Client report
WLG2010-31, Prepared for the Ministry of Agriculture and Forestry (MAF) by
the National Institute of Water and Atmospheric Research (NIWA), MAF,
Wellington, New Zealand, 80 pp.
1430
Chapter 25 Australasia
25
Munday, P.L., N.E. Crawley, and G.E. Nilsson, 2009: Interacting effects of elevated
temperature and ocean acidification on the aerobic performance of coral reef
fishes. Marine Ecology Progress Series, 388, 235-242.
Murphy, H., A. Liedloff, R.J. Williams, K.J. Williams, and M. Dunlop, 2012: Queensland’s
Biodiversity under Climate Change: Terrestrial Ecosystems. CSIRO Climate
Adaptation Flagship Working Paper No. 12C, Commonwealth Scientific and
Industrial Research Organization (CSIRO), Canberra, ACT, Australia, 110 pp.
Mustelin, J., N. Kuruppu, A.M. Kramer, J. Daron, K. de Bruin, and A.G. Noriega, 2013:
Climate adaptation research for the next generation. Climate and Development,
5(3), 189-193.
Naish, S., K. Mengersen, W. Hu, and S. Tong, 2013: Forecasting the future risk of
Barmah Forest Virus disease under climate change scenarios in Queensland,
Australia. PLoS ONE, 8(5), e62843, doi:10.1371/journal.pone.0062843.
Nana, G., F. Stokes, and W. Molano, 2011a: The Asset Base, Income, Expenditure and
GDP of the 2010 Māori Economy. Report for the Māori Economic Taskforce by
Business and Economic Research Ltd. (BERL Economics), Māori Economic
Taskforce Secretariat, Wellington, New Zealand, 40 pp.
Nana, G., F. Stokes, and W. Molano, 2011b: The Māori Economy, Science and
Innovation Scenarios. Report for the Māori Economic Taskforce by Business and
Economic Research Ltd. (BERL Economics), ori Economic Taskforce Secretariat,
Wellington, New Zealand, 40 pp.
NDIR, 2011: Natural Disaster Insurance Review: Inquiry into Flood Insurance and
Related Matters. Natural Disaster Insurance Review (NDIR), c/o The Treasury,
Parkes, ACT, Australia, 149 pp.
Nearing, M.A., F.F. Pruski, and M.R. O’Neal, 2004: Expected climate change impacts
on soil erosion rates: a review. Journal of Soil and Water Conservation, 59(1),
43-50.
Nelson, R., M. Howden, and M.S. Smith, 2008: Using adaptive governance to rethink
the way science supports Australian drought policy. Environmental Science &
Policy, 11(7), 588-601.
Nelson, R., P. Kokic, S. Crimp, P. Martin, H. Meinke, S.M. Howden, P. de Voil, and U.
Nidumolu, 2010: The vulnerability of Australian rural communities to climate
variability and change: Part II integrating impacts with adaptive capacity.
Environmental Science and Policy, 13(1), 18-27.
Neuheimer, A.B., R.E. Thresher, J.M. Lyle, and J.M. Semmens, 2011: Tolerance limit
for fish growth exceeded by warming waters. Nature Climate Change, 1(2),
110-113.
Newton, P.C.D., V. Allard, R.A. Carran, and M. Lieffering, 2006: Impacts of elevated
CO
2
on a grassland grazed by sheep: the New Zealand FACE experiment. In:
Managed Ecosystems and CO
2
: Case Studies, Processes, and Perspectives
[Nosberger, J., S.P. Long, R.J. Norby, M. Stitt, G.R. Hendrey, and H. Blum (eds.)].
Springer-Verlag, Berlin, Germany, pp. 157-171.
Newton, P.C.D., M. Lieffering, A.J. Parsons, S.C. Brock, P.W. Theobald, C.L. Hunt, D.
Luo, and M.J. Hovenden, 2014: Selective grazing modifies previously anticipated
responses of plant community composition to elevated CO
2
in a temperate
grassland. Global Change Biology, 20(1), 158-169.
NFRAG, 2008: National Flood Risk Advisory Group (NFRAG): flood risk management
in Australia. Australian Journal of Emergency Management, 23, 21-27.
Nguyen, M., X. Wang, and D. Chen, 2010: An Investigation of Extreme Heatwave
Events and their Effects on Building and Infrastructure. Commonwealth
Scientific and Industrial Research Organization (CSIRO), Canberra, ACT,
Australia, 112 pp.
Nicholls, N., 2006: Detecting and attributing Australian climate change: a review.
Australian Meteorological Magazine, 55(3), 199-211.
Nicholls, N., 2010: Local and remote causes of the southern Australian autumn-
winter rainfall decline, 1958-2007. Climate Dynamics, 34(6), 835-845.
Nicholls, N., 2011: Comments on “Influence of location, population, and climate on
building damage and fatalities due to Australian bushfire: 1925-2009”.
Weather, Climate, and Society, 3(1), 61-62.
Nicholls, N. and D. Collins, 2006: Observed climate change in Australia over the past
century. Energy & Environment, 17(1), 1-12.
Nicholls, N., P. Uotila, and L. Alexander, 2010: Synoptic influences on seasonal,
interannual and decadal temperature variations in Melbourne, Australia.
International Journal of Climatology, 30(9), 1372-1381.
Nidumolu, U., S. Crimp, D. Gobbett, A. Laing, M. Howden, and S. Little, 2011: Heat
Stress in Dairy Cattle in Northern Victoria: Responses to a Changing Climate.
CSIRO Climate Adaptation Flagship Working Paper No. 10, Commonwealth
Scientific and Industrial Research Organization (CSIRO), Canberra, ACT,
Australia, 63 pp.
Nilsson, G.E., D.L. Dixson, P. Domenici, M.I. McCormick, C. Sorensen, S.A. Watson,
and P.L. Munday, 2012: Near-future carbon dioxide levels alter fish behaviour
by interfering with neurotransmitter function. Nature Climate Change, 2(3),
201-204.
NIWA, MWH, GNS Science, and BRANZ, 2012: Impacts of Climate Change on Urban
Infrastructure and the Built Environment: Toolbox Handbook. National Institute
of Water and Atmospheric Research (NIWA), MWH New Zealand Ltd. (MWH),
GNS Science, and Building Research Association of New Zealand (BRANZ),
NIWA, Wellington, New Zealand, 32 pp.
NNTR, 2013: National Native Title Register (NNTR) Determinations of Native Title –
Boundaries and Core Attributes. National Native Title Tribunal (NNTT) Dataset
as per s192 Native Title Act 1993 (Commonwealth), Geoscience Australia,
Canberra, ACT, Australia, online dataset, accessed 31 August 2013, at
www.ga.gov.au/meta/ANZCW0703011416.html.
Norman, B., 2009: Principles for an intergovernmental agreement for coastal
planning and climate change in Australia. Habitat International, 33(3), 293-
299.
Norman, B., 2010: A Low Carbon and Resilient Urban Future: An Integrated Approach
to Planning for Climate Change. Technical Report prepared for the Department
of Climate Change and Energy Efficiency (DCCEE), DCCEE, Canberra, ACT,
Australia, 60 pp.
Norman, B., W. Steffen, R. Webb, A. Capon, W. Maher, C. Woodroffe, K. Rogers, R.
Tanton, Y. Vidyattama, J. Lavis, H. Sinclair, and B. Weir, 2013: South East Coastal
Adaptation (SECA): Coastal Urban Climate Futures in SE Australia from
Wollongong to Lakes Entrance: Final Report. National Climate Change
Adaptation Research Facility (NCCARF), Griffith University, Gold Coast Campus,
Southport, QLD, Australia, 130 pp.
Norman-Lopez, A., S. Pascoe, and A.J. Hobday, 2011: Potential economic impacts of
climate change on Australian fisheries and the need for adaptive management.
Climate Change Economics, 2, 209-235.
NRC, 2012: Framework for Assessing and Recommending Upgraded Catchment
Action Plans: Version 2.0. Natural Resources Commission (NRC), New South
Wales Government, Sydney, NSW, Australia, 27 pp.
NRMMC, 2010: National Feral Camel Action Plan: A National Strategy for the
Management of Feral Camels in Australia. Vertebrate Pests Committee,
Natural Resource Management Ministerial Council (NRMMC), Department of
Sustainability, Environment, Water, Population and Communities, Canberra,
ACT, Australia, 68 pp.
Nursey-Bray, M., D. Fergie, V. Arbon, L. Rigney, R. Palmer, J. Tibby, N. Harvey, and L.
Hackworth, 2013: Community Based Adaptation to Climate Change: The
Arabana, South Australia: Final Report. National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 133 pp.
NWC, 2009: Australian Water Reform 2009: Second Biennial Assessment of Progress
in Implementation of the National Water Initiative. National Water Commission
(NWC), Canberra, ACT, Australia, 288 pp.
NWC, 2010: The Impacts of Water Trading in the Southern Murray-Darling Basin: An
Economic, Social and Environmental Assessment. National Water Commission
(NWC), Canberra, ACT, Australia, 238 pp.
NWC, 2011: The National Water Initiative – Securing Australia’s Water Future: 2011
Assessment. National Water Commission (NWC), Canberra, ACT, Australia,
354 pp.
NZIER, 2003: Māori Economic Development: Te Ohanga Whakaketanga Māori. New
Zealand Institute of Economic Research (NZIER), Inc., Wellington, New Zealand,
106 pp.
O’Connell, M. and R. Hargreaves, 2004: Climate Change Adaptation: Guidance on
Adapting New Zealand’s Built Environment for the Impacts of Climate Change.
Study Report No.130, Building Research Association of New Zealand (BRANZ),
Wellington, New Zealand, 44 pp.
O’Donnell, L., 2007: Climate Change: An Analysis of the Policy Considerations for
Climate Change for the Review of the Canterbury Regional Policy Statement.
Report R07/4, Environment Canterbury, Christchurch, New Zealand, 75 pp.
O’Leary, G., B. Christy, A. Weeks, C. Beverly, J. Nuttall, P. Riffkin, G. Fitzgerald, and R.
Norton, 2010: Modelling the growth and grain yield response of wheat crops
under Free Air Carbon dioxide Enrichment at Horsham and throughout Victoria,
Australia. In: “Food Security from Sustainable Agriculture,” 15th Agronomy
Conference, 15-18 November 2010, Lincoln, New Zealand [Dove, H. and R.A.
Culvenor (eds.)]. Australian Society of Agronomy, Gosford, NSW, Australia,
www.regional.org.au/au/asa/2010/climate-change/co2/7013_olearyg.htm.
1431
Australasia Chapter 25
25
O’Neill, S.J. and J. Handmer, 2012: Responding to bushfire risk: the need for
transformative adaptation. Environmental Research Letters, 7(1), 014018,
doi:10.1088/1748-9326/7/1/014018.
Odeh, I.O.A., D.K.Y. Tan, and T. Ancev, 2011: Potential suitability and viability of
selected biodiesel crops in Australian marginal agricultural lands under current
and future climates. Bioenergy Research, 4(3), 165-179.
OECD, 2010: Sustainable Management of Water Resources in Agriculture. Organisation
for Economic Co-Operation and Development (OECD), OECD Publishing, Paris,
France,118 pp.
OECD, 2011: Dairy. In: OECD-FAO Agricultural Outlook 2011 [Organisation for
Economic Co-Operation and Development (OECD) and Food and Agriculture
Organization of the United Nations (FAO) (eds.)]. OECD Publishing, Paris, France,
pp. 159-173.
OECD, 2013a: OECD Factbook 2013: Economic, Environmental and Social Statistics.
Organisation for Economic Co-Operation and Development (OECD), OECD
Publishing, Paris, France, 292 pp.
OECD, 2013b: Water Security for Better Lives. OECD Studies on Water, Organisation
for Economic Co-Operation and Development (OECD), OECD Publishing, Paris,
France, 172 pp.
Ogden, J., L. Basher, and M.S. McGlone, 1998: Fire, forest regeneration and links with
early human habitation: evidence from New Zealand. Annals of Botany, 81,
687-696.
Oliver, R. and I. Webster, 2011: Water for the environment. In: Water: Science and
Solutions for Australia [Prosser, I.P. (ed.)]. Commonwealth Scientific and
Industrial Research Organization (CSIRO) Publishing, Collingwood, VIC,
Australia, pp. 119-134.
Oxford Economics, 2009: Valuing the Effects of Great Barrier Reef Bleaching.
Prepared for the Great Barrier Reef Foundation by Oxford Economics, Great
Barrier Reef Foundation, Newstead, QLD, Australia, 95 pp.
Packman, D., D. Ponter, and T. Tutua-Nathan, 2001: Climate Change Working Paper:
Māori Issues. The Department of the Prime Minister and Cabinet, New Zealand
Climate Change Programme, Wellington, New Zealand, 18 pp.
Palutikof, J., M. Parry, M. Stafford-Smith, A.J. Ash, S.L. Boulter, and M. Waschka, 2013:
The past, present and future of adaptation: setting the context and naming the
challenges. In: Climate Adaptation Futures [Palutikof, J., S.L. Boulter, A.J. Ash,
M. Stafford-Smith, M. Parry, M. Waschka, and D. Guitart (eds.)]. John Wiley &
Sons, Brisbane, QLD, Australia, pp. 3-29.
Park, S.E., S. Crimp, N.G. Inman-Bamber, and Y.L. Everingham, 2010: Sugarcane. In:
Adapting Agriculture to Climate Change. Preparing Australian Agriculture,
Forestry and Fisheries for the Future [Stokes, C. and M. Howden (eds.)].
Commonwealth Scientific and Industrial Research Organization (CSIRO)
Publishing, Collingwood, VIC, Australia, pp. 85-99.
Park, S.E., S. Crimp, A.M. Dowd, P. Ashworth, and E. Mendham, 2011: Understanding
Climate Change Adaptation as a National Challenge. Climate Adaptation
Flagship Working Paper No. 8, Commonwealth Scientific and Industrial
Research Organization (CSIRO) Climate Adaptation Flagship, Canberra, ACT,
Australia, 24 pp.
Park, S.E., N.A. Marshall, E. Jakku, A.M. Dowd, S.M. Howden, E. Mendham, and A.
Fleming, 2012: Informing adaptation responses to climate change through
theories of transformation. Global Environmental Change, 22(1), 115-126.
Parker, D., C. Folland, A. Scaife, J. Knight, A. Colman, P. Baines, and B. Dong, 2007:
Decadal to multidecadal variability and the climate change background. Journal
of Geophysical Research: Atmospheres, 112(D18), D18115, doi:10.1029/
2007JD008411.
Parsons Brinkerhoff, 2009: Energy Network Infrastructure and the Climate Change
Challenge. ENA Briefing Note, Report by Parsons Brinkerhoff (PB) Strategic
Consulting Group in the Australia-Pacific Region for the Energy Networks
Association (ENA) Australia, ENA Australia, Barton, ACT, Australia, 131 pp.
Peace, R., 2001: Social exclusion: a concept in need of definition? Social Policy Journal
of New Zealand, 16, 17-36.
Pearce, A. and M. Feng, 2007: Observations of warming on the Western Australian
continental shelf. Marine and Freshwater Research, 58(10), 914-920.
Pearce, H.G., G. Cameron, S.A.J. Anderson, and M. Dudfield, 2008: An overview of
fire management in New Zealand forestry. New Zealand Journal of Forestry,
53(3), 7-11.
Pearce, H.G., J. Kerr, A. Clark, B. Mullan, D. Ackerley, T. Carey-Smith, and E. Yang, 2011:
Improved Estimates of the Effect of Climate Change on NZ Fire Danger. MAF
Technical Paper No. 2011/13, Ministry of Agriculture and Forestry, Wellington,
New Zealand, 84 pp.
Pearce, M., E. Willis, and T. Jenkin, 2007: Aboriginal people’s attitudes towards paying
for water in a water-scarce region of Australia. Environment Development and
Sustainability, 9(1), 21-32.
Pearman, G., 2009: Climate change and security: why is this on the agenda? In:
Climate Change and Security: Planning for the Future [Boston, J., P. Nel, and
M. Righarts (eds.)]. Institute of Policy Studies, Victoria University of Wellington,
Wellington, New Zealand, pp. 9-36.
Peel, M.C., T.A. McMahon, and B.L. Finlayson, 2004: Continental differences in the
variability of annual runoff-update and reassessment. Journal of Hydrology,
295(1-4), 185-197.
Pecl, G., S. Frusher, C. Gardner, M. Haward, A. Hobday, S. Jennings, M. Nursey-Bray,
A. Punt, H. Revill, and I. vanPutten, 2009: The East Coast Tasmanian Rock
Lobster Fishery Vulnerability to Climate Change Impacts and Adaptation
Response Options. Report to the Department of Climate Change (DCC), DCC,
Canberra, ACT, Australia, 100 pp.
Penny, G. and D. King, 2010: Transfer of Renewable Energy Technology Systems to
Rural Māori Communities. NIWA Client Report No. AKL2007-051, Prepared for
the Foundation for Research, Science and Technology (FRST) by the National
Institute of Water and Atmospheric Research (NIWA), FRST, Wellington, New
Zealand, 59 pp.
Petheram, C., P. Rustomji, F. Chiew, W. Cai, T. McVicar, J. Vleeshouwer, T. Van Neil, L.
Li, R. Donohue, R. Cresswell, J. Teng, and J.-M. Perraud, 2012: Estimating the
impact of projected climate change on runoff across the tropical savannas and
semi-arid rangelands of northern Australia. Journal of Hydrometeorology,
13(2), 483-503.
Petheram, L., K.K. Zander, B.M. Campbell, C. High, and N. Stacey, 2010:Strange
changes’: Indigenous perspectives of climate change and adaptation in NE
Arnhem Land (Australia). Global Environmental Change, 20(4), 681-692.
Petrone, K., J. Hughes, T. Van Niel, and R. Silberstein, 2010: Streamflow decline in
southwestern Australia, 1950-2008. Geophysical Research Letters, 37, L043102,
doi:10.1029/2010GL043102.
Pettit, C.J., C.M. Raymond, B.A. Bryan, and H. Lewis, 2011: Identifying strengths and
weaknesses of landscape visualisation for effective communication of future
alternatives. Landscape and Urban Planning, 100(3), 231-241.
Pham, T.D., D.G. Simmons, and R. Spurr, 2010: Climate change-induced economic
impacts on tourism destinations: the case of Australia. Journal of Sustainable
Tourism, 18(3), 449-473.
Phelan, L., 2011: Managing climate risk: extreme weather events and the future of
insurance in a climate-changed world. Australian Journal of Environmental
Management, 18, 223-232.
Pickering, C.M. and R.C. Buckley, 2010: Climate response by the ski industry: the
shortcomings of snowmaking for Australian resorts. AMBIO: A Journal of the
Human Environment, 39(5-6), 430-438.
Pickering, C.M., W. Hill, and K. Green, 2008: Vascular plant diversity and climate
change in the alpine zone of the Snowy Mountains, Australia. Biodiversity and
Conservation, 17(7), 1627-1644.
Pickering, C.M., J.G. Castley, and M. Burtt, 2010: Skiing less often in a warmer world:
attitudes of tourists to climate change in an Australian ski resort. Geographical
Research, 48(2), 137-147.
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.
Pittock, J., 2013: Lessons from adaptation to sustain freshwater environments in the
Murray-Darling Basin, Australia. Wiley Interdisciplinary Reviews: Climate
Change, 4(5), 429-438.
Pittock, J. and C.M. Finlayson, 2011: Australia’s Murray-Darling Basin: freshwater
ecosystem conservation options in an era of climate change. Marine and
Freshwater Research, 62(3), 232-243.
Pittock, J., L.J. Hansen, and R. Abell, 2008: Running dry: freshwater biodiversity,
protected areas and climate change. Biodiversity, 9(3-4), 30-38.
PMSEIC, 2010a: Australia and Food Security in a Changing World: Can We Feed
Ourselves and Help Feed the World in the Future? Report prepared by the
independent PMSEIC Expert Working Group on Food Security for the Prime
Minister’s Science, Engineering and Innovation Council (PMSEIC), PMSEIC,
Canberra, ACT, Australia, 80 pp.
PMSEIC, 2010b: Challenges at Energy-Water-Carbon Intersections. Report prepared
by the independent PMSEIC Expert Working Group on Food Security for the
Prime Minister’s Science, Engineering and Innovation Council (PMSEIC),
PMSEIC, Canberra, ACT, Australia, 76 pp.
1432
Chapter 25 Australasia
25
Polglase, P., A. Reeson, C. Hawkins, K. Paul, A. Siggins, J. Turner, D. Crawford, T.
Jovanovic, T. Hobbs, K. Opie, J. Carwardine, and A. Almeida, 2011: Opportunities
for Carbon Forestry in Australia: Economic Assessment and Constraints to
Implementation. Commonwealth Scientific and Industrial Research Organization
(CSIRO), Canberra, ACT, Australia, 24 pp.
Poloczanska, E.S., R.C. Babcock, A. Butler, A. Hobday, O. Hoegh-Guldberg, T.J. Kunz,
R. Matear, D.A. Milton, T.A. Okey, and A.J. Richardson, 2007: Climate change
and Australian marine life. Oceanography and Marine Biology, 45, 407-478.
Pomeroy, A., 1996: Relocating policies for sustainable agriculture under the umbrella
of rural development in Australia, USA and New Zealand. New Zealand
Geographer, 52(2), 84-87.
Poot, J., 2009: Trans-Tasman Migration, Transnationalism and Economic Development
in Australasia. Motu Working Paper 09-05, Motu Economic and Public Policy
Research, Wellington, New Zealand, 21 pp.
Post, D.A., J. Teng, F. Chiew, B. Wang, J. Vaze, S. Marvanek, S.W. Franks, E. Boegh, E.
Blyth, D.M. Hannah, and K.K. Yilmaz, 2011: Non-linearity of the runoff response
across southeastern Australia to increases in global average temperature. In:
Hydro-Climatology: Variability and Change [Franks, S.W., E. Boegh, E. Blyth,
D.M. Hannah, and K.K. Yilmaz (eds.)]. IAHS Publication 344, International
Association of Hydrological Sciences (IAHS), Oxfordshire, UK, pp. 188-194.
Post, D.A., F. Chiew, J. Teng, N.R. Viney, F.L.N. Ling, G. Harrington, R.S. Crosbie, B. Graham,
S. Marvanek, and R. McLoughlin, 2012: A robust methodology for conducting
large-scale assessments of current and future water availability and use: a case
study in Tasmania, Australia. Journal of Hydrology, 412-413, 233-245.
Potter, N.J. and F. Chiew, 2011: An investigation into changes in climate characteristics
causing the recent very low runoff in the southern Murray-Darling Basin using
rainfall-runoff models. Water Resources Research, 47, W00G10, doi:10.1029/
2010WR010333.
Potter, N.J., F. Chiew, and A.J. Frost, 2010: An assessment of the severity of recent
reductions in rainfall and runoff in the Murray-Darling Basin. Journal of
Hydrology, 381(1-2), 52-64.
Power, S., F. Tseitkin, S. Torok, B. Lavery, R. Dahni, and B. McAvaney, 1998: Australian
temperature, Australian rainfall and the Southern Oscillation, 1910-1992:
coherent variability and recent changes. Australian Meteorological Magazine,
47(2), 85-101.
Power, S., B. Sadler, and N. Nicholls, 2005: The influence of climate science on water
management in Western Australia: lessons for climate scientists. Bulletin of the
American Meteorological Society, 86(6), 839-844.
Poyck, S., J. Hendrikx, H. McMillan, E. Hreinsson, and R. Woods, 2011: Combined
snow and streamflow modelling to estimate impacts of climate change on
water resources in the Clutha River, New Zealand. Journal of Hydrology, 50(2),
293-312.
Pratchett, M.S., L.K. Bay, P.C. Gehrke, J.D. Koehn, K. Osborne, R.L. Pressey, H.P.A.
Sweatman, and D. Wachenfeld, 2011: Contribution of climate change to
degradation and loss of critical fish habitats in Australian marine and freshwater
environments. Marine and Freshwater Research, 62(9), 1062-1081.
Preston, B. and R.N. Jones, 2008: Screening climatic and non-climatic risks to
Australian catchments. Geographical Research, 46(3), 258-274.
Preston, B. and M. Stafford-Smith, 2009: Framing Vulnerability and Adaptive Capacity
Assessment: Discussion Paper. CSIRO Climate Adaptation Flagship Working
Paper No. 2, Commonwealth Scientific and Industrial Research Organization
(CSIRO), Canberra, ACT, Australia, 52 pp.
Preston, B., T.F. Smith, C. Brooke, R. Gorddard, T. Measham, G. Withycombe, B.
Beveridge, C. Morrison, K. McInnes, and D. Abbs, 2008: Mapping Climate
Change Vulnerability in the Sydney Coastal Councils Group. Report prepared
by the Commonwealth Scientific and Industrial Research Organization (CSIRO)
for the Sydney Coastal Councils Group (SCCG) and Australian Government
Department of Climate Change (DCC), SCCG Inc., Sydney, NSW, Australia, 117 pp.
Preston, B., C. Brooke, T. Measham, T. Smith, and R. Gorddard, 2009: Igniting change
in local government: lessons learned from a bushfire vulnerability assessment.
Mitigation and Adaptation Strategies for Global Change, 14(3), 251-283.
Preston, B., R. Westaway, and E. Yuen, 2011: Climate adaptation planning in practice:
an evaluation of adaptation plans from three developed nations. Mitigation
and Adaptation Strategies for Global Change, 16(4), 407-438.
Pride, S., B. Frame, and D. Gill, 2010: Futures literacy in New Zealand. Journal of
Futures Studies, 15(1), 135-144.
Prober, S.M. and M. Dunlop, 2011: Climate change: a cause for new biodiversity
conservation objectives but let’s not throw the baby out with the bathwater.
Ecological Management and Restoration, 12, 2-3.
Productivity Commission, 2009: Government Drought Support. Report No. 46, Final
Inquiry Report, Productivity Commission, Melbourne Office, Melbourne, VIC,
Australia, 431 pp.
Productivity Commission, 2012: Barriers to Effective Climate Change Adaptation.
Inquiry Report No. 59, Final Inquiry Report, Productivity Commission, Canberra
Office, Canberra, ACT, Australia, 385 pp.
Prosser, I.P., I.D. Rutherfurd, J.M. Olley, W.J. Young, P.J. Wallbrink, and C.J. Moran,
2001: Large-scale patterns of erosion and sediment transport in rivers networks,
with examples from Australia. Freshwater and Marine Research, 52(1), 81-99.
Prowse, T.A.A. and B.W. Brook, 2011: Climate change, variability and conservation
impacts in Australia. Pacific Conservation Biology, 17(3), 168-178.
Purdie, H., A. Mackintosh, W. Lawson, B. Anderson, U. Morgenstern, T. Chinn, and P.
Mayewski, 2011: Interannual variability in net accumulation on Tasman Glacier
and its relationship with climate. Global and Planetary Change, 77(3-4), 142-
152.
PwC, 2011: Economic Impact of Queensland’s Natural Disasters. March 2011.
PricewaterhouseCoopers (PwC), Sydney, NSW, Australia, 5 pp.
QFCI, 2012: Queensland Floods Commission of Inquiry: Final Report. Queensland
Floods Commission of Inquiry (QFCI), Brisbane, QLD, Australia, 654 pp.
QRC, 2011: Queensland Resources Council Submission to Queensland Floods
Commission of Inquiry. Queensland Resources Council (QRC), Brisbane, QLD,
Australia, 43 pp.
Queensland Government, 2011: Understanding Floods: Questions and Answers.
Queensland Government, Brisbane, QLD, Australia, 36 pp.
Quiggin, J., D. Adamson, P. Schrobback, and S. Chambers, 2008: The Implications for
Irrigation in the Murray-Darling Basin. Paper prepared for the Garnaut Climate
Change Review, Canberra, ACT, Australia, 59 pp.
Quiggin, J., D. Adamson, S. Chambers, and P. Schrobback, 2010: Climate change,
uncertainty, and adaptation: the case of irrigated agriculture in the Murray-
Darling Basin in Australia. Canadian Journal of Agricultural Economics/Revue
canadienne d’agroeconomie, 58(4), 531-554.
Qureshi, E.M., M.A. Hanjra, and J. Ward, 2013a: Impact of water scarcity in Australia
on global food security in an era of climate change. Food Policy, 38(0), 136-
145.
Qureshi, E.M., S.M. Whitten, M. Mohammed, S. Marvanek, and A. Elmahdi, 2013b: A
biophysical and economic model of agriculture and water in the Murray-Darling
Basin, Australia. Environmental Modelling & Software, 41, 98-106.
QUT, 2010: Impacts and Adaptation Response of Infrastructure and Communities to
Heatwaves: The Southern Australian Experience of 2009. CRICOS No. 00213J,
Prepared for the National Climate Change Adaptation Research Facility
(NCCARF) by the Institute for Sustainable Resources, Queensland University of
Technology (QUT), Monash University and the University of Southern Queensland
with contributions from Access Macquaire Ltd. and the Royal Melbourne Insititute
of Technology, NCCARF, Griffith University, Gold Coast Campus, Southport, QLD,
Australia, 153 pp.
Radcliffe, J.E. and J.A. Baars, 1987: Temperate pasture production. In: Ecosystems of
the World [Snaydon, R.W. (ed.)]. Elsevier, Amsterdam, Netherlands, pp. 7-17.
Rafter, A. and D.J. Abbs, 2009: An analysis of future changes in extreme rainfall over
Australian regions based on GCM simulations and Extreme Value Analysis.
CAWCR Research Letters (The Centre for Australian Weather and Climate
Research), 3, 44-49, www.cawcr.gov.au/publications/researchletters/CAWCR_
Research_Letters_3.pdf.
Rajanayaka, C., J. Donaggio, and H. McEwan, 2010: Update of Water Allocation Data
and Estimate of Actual Water Use of Consented Takes 2009-10. Report No.
H10002/3, October 2010, Prepared for the Ministry for the Environment (MfE)
by Aqualinc Research Ltd., MfE, Wellington, New Zealand, 118 pp.
Randall, A., T. Capon, T. Sanderson, D. Merrett, and G. Hertzler, 2012: Making
Decisions under the Risks and Uncertainties of Future Climates. Final report for
the National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 66 pp.
Rau, G.H., E.L. McLeod, and O. Hoegh-Guldberg, 2012: The need for new ocean
conservation strategies in a high-carbon dioxide world. Nature Climate Change,
2(10), 720-724.
Raymond, C.M. and G. Brown, 2011: Assessing spatial associations between
perceptions of landscape value and climate change risk for use in climate
change planning. Climatic Change, 104(3-4), 653-678.
Raymond, C.M. and J. Spoehr, 2013: The acceptability of climate change in agricultural
communities: comparing responses across variability and change. Journal of
Environmental Management, 115, 69-77.
1433
Australasia Chapter 25
25
Raymond, C.M., G. Brown, and G.M. Robinson, 2011: The influence of place
attachment, and moral and normative concerns on the conservation of native
vegetation: a test of two behavioural models. Journal of Environmental
Psychology, 31(4), 323-335.
RBA, 2006: Statement on Monetary Policy November 2006. Reserve Bank of
Australia (RBA), Canberra, ACT, Australia, 59 pp.
RBA, 2011: Box B: an update on the impact of natural disasters in Queensland. In:
Statement on Monetary Policy – May 2011. Reserve Bank of Australia (RBA),
Canberra, ACT, Australia, pp. 40-44.
Reijnders, L. and M.A.J. Huijbregts, 2007: Life cycle greenhouse gas emissions, fossil
fuel demand and solar energy conversion efficiency in European bioethanol
production for automotive purposes. Journal of Cleaner Production, 15(18),
1806-1812.
Reisinger, A., B. Mullan, M.R. Manning, D. Wratt, and R. Nottage, 2010: Global and
local climate change scenarios to support adaptation in New Zealand. In:
Climate Change Adaptation in New Zealand: Future Scenarios and Some
Sectoral Perspectives [Nottage, R., D. Wratt, J. Bornman, and K. Jones
(eds.)]. New Zealand Climate Change Centre, Wellington, New Zealand,
pp. 26-43.
Reisinger, A., D. Wratt, S. Allan, and H. Larsen, 2011: The role of local government in
adapting to climate change: lessons from New Zealand. In: Climate Change
Adaptation in Developed Nations: From Theory to Practice [Ford, J.D. and L.
Berrang-Ford (eds.)]. Advances in Global Change Research Vol. 42, Springer
Science, Dordrecht, Netherlands, pp. 303-319.
Reisinger, A., P. Havlik, K. Riahi, O. Vliet, M. Obersteiner, and M. Herrero, 2012:
Implications of alternative metrics for global mitigation costs and greenhouse
gas emissions from agriculture. Climatic Change, 117(4), 677-690.
Reisinger, A., J. Lawrence, G. Hart, and R. Chapman, 2013: From coping to resilience:
the role of managed retreat in highly developed coastal regions. In: Climate
Change and the Coast: Building Resilient Communities [Glavovic, B., R. Kaye, M.
Kelly, and A. Travers (eds.)]. CRC Press, London, UK (in press), www.crcpress.com/
product/isbn/9780415464871.
Ren, Z., X. Wang, and Z. Chen, 2012: Energy efficient houses for heat stress mitigation:
2009 Melbourne heatwave scenarios. In: 10th International Conference on
Healthy Buildings 2012, 8-12 July 2012, Brisbane, QLD, Australia. Session 4D:
Sustainability and Green Buildings – Construction, Design and Material: 4D.5,
Sponsored by the International Society of Indoor Air Quality and Climate
(ISIAQ), Brisbane, QLD, Australia.
Renwick, J., P. Mladenov, J. Purdie, A. McKerchar, and D. Jamieson, 2009: The effects
of climate variability and change upon renewable electricity in New Zealand.
In: Climate Change Adaptation in New Zealand: Future Scenarios and Some
Sectoral Perspectives [Nottage, R., D. Wratt, J. Bornman, and K. Jones (eds.)].
New Zealand Climate Change Centre, Wellington, New Zealand, pp. 70-81.
Reser, J.P. and J.K. Swim, 2011: Adapting to and coping with the perceived threat
and unfolding impacts of global climate change. American Psychologist, 66(4),
277-289.
Reser, J.P., S.A. Morrissey, and M. Ellul, 2011: The threat of climate change:
psychological response, adaptation, and impacts. In: Climate Change and
Human Well Being: Global Challenges and Opportunities [Weissbecker, I. (ed.)].
International and Cultural Psychology Series, Springer, New York, NY, USA, pp.
19-42.
Reser, J.P., G.L. Bradley, and M.C. Ellul, 2012a: Coping with climate change: bringing
psychological adaptation in from the cold. In: Handbook of the Psychology of
Coping: Psychology of Emotions, Motivations and Actions [Molinelli, B. and V.
Grimaldo (eds.)]. Nova Science Publishers, New York, NY, USA, pp. 1-34.
Reser, J.P., G.L. Bradley, A.I. Glendon, M.C. Ellul, and R. Callaghan, 2012b: Public Risk
Perceptions, Understandings and Responses to Climate Change and Natural
Disasters in Australia, 2010 and 2011: Final Report. Report prepared by Griffith
University, School of Psychology, Behavioural Basis of Health, Griffith Climate
Change Response Program (GCCRP), National Climate Change Adaptation
Research Facility (NCCARF), Griffith University, Gold Coast Campus, Southport,
QLD, Australia, 245 pp.
Reser, J.P., G.L. Bradley, A.I. Glendon, M.C. Ellul, and R. Callaghan, 2012c: Public Risk
Perceptions, Understandings, and Responses to Climate Change and Natural
Disasters in Australia and Great Britain. Report prepared by Griffith University,
School of Psychology, Behavioural Basis of Health, Griffith Climate Change
Response Program (GCCRP), National Climate Change Adaptation Research
Facility (NCCARF), Griffith University, Gold Coast Campus, Southport, QLD,
Australia, 274 pp.
Reside, A.E., J. Van Der Wal, B.L. Phillips, L.P. Shoo, D.F. Rosauer, B.J. Anderson, J.A.
Welbergen, C. Moritz, S. Ferrier, T.D. Harwood, K.J. Williams, Brendan Mackey,
S. Hugh, and S.E. Williams, 2013: Climate Change Refugia for Terrestrial
Biodiversity: Defining Areas that Promote Species Persistence and Ecosystem
Resilience in the Face of Global Climate Change: Final Report. Prepared by
James Cook University, National Climate Change Adaptation Research Facility
(NCCARF), Griffith University, Gold Coast Campus, Southport, QLD, Australia,
216 pp.
Rhodes, B.G., J.E. Fagan, and K.S. Tan, 2012: Responding to a rapid climate shift
experiences from Melbourne, Australia. In: IWA World Congress on Water, Climate
and Energy, 13-18 May 2012, Dublin, Ireland. Organized by the International
Water Association (IWA), Dublin, Ireland, keynote.conference-services.net/
resources/444/2653/pdf/IWAWCE2012_0222.pdf.
Richardson, A.J. and E.S. Poloczanska, 2009: Australia’s oceans. In: A Marine Climate
Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska,
E.S., A.J. Hobday, and A.J. Richardson (eds.)]. NCCARF Publication 05/09,
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 6 pp.
Richardson, A.J., D. McKinnon, and K.M. Swadling, 2009: Zooplankton. In: A Marine
Climate Change Impacts and Adaptation Report Card for Australia 2009
[Poloczanska, E.S., A.J. Hobday, and A.J. Richardson (eds.)]. NCCARF Publication
05/09, National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 15 pp.
Rickards, L. and S.M. Howden, 2012: Transformational adaptation: agriculture and
climate change. Crop and Pasture Science, 63(3), 240-250.
Ridgway, K.R., 2007: Long-term trend and decadal variability of the southward
penetration of the East Australian Current. Geophysical Research Letters,
34(13), L030393, doi:10.1029/2007GL030393.
Risbey, J.S., M.J. Pook, P.C. McIntosh, M.C. Wheeler, and H.H. Hendon, 2009: On the
remote drivers of rainfall variability in Australia. Monthly Weather Review,
137(10), 3233-3253.
Ritchie, E.G. and E.E. Bolitho, 2008: Australia’s savanna herbivores: bioclimatic
distributions and an assessment of the potential impact of regional climate
change. Physiological and Biochemical Zoology, 81(6), 880-890.
Rive, V. and T. Weeks, 2011: Adaptation to climate change in New Zealand. In: Climate
Change Law and Policy in New Zealand [Cameron, A. (ed.)]. LexisNexis,
Wellington, New Zealand, pp. 345-392.
Roberts, D.A., E.L. Johnston, and N.A. Knott, 2010: Impacts of desalination plant
discharges on the marine environment: a critical review of published studies.
Water Research, 44(18), 5117-5128.
Robertson, M., 2010: Agricultural productivity in Australia and New Zealand: trends,
constraints and opportunities. In: Food Security from Sustainable Agriculture.”
15
th
Agronomy Conference, 15-18 November 2010, Lincoln, New Zealand [Dove,
H. and R.A. Culvenor (eds.)]. Organized by the Australian Society of Agronomy,
Gosford, NSW, Australia, www.regional.org.au/au/asa/2010/plenary/climate-
change/7402_robertsonmj.htm.
Rogan, R., M. O’Connor, and P. Horwitz, 2005: Nowhere to hide: awareness and
perceptions of environmental change, and their influence on relationships with
place. Journal of Environmental Psychology, 25(2), 147-158.
Rogers, K., N. Saintilan, and C. Copeland, 2012: Modelling wetland surface elevation
dynamics and its application to forecasting the effects of sea-level rise on
estuarine wetlands. Ecological Modelling, 244, 148-157.
Roiko, A., R.B. Mangoyana, S. McFallan, R.W. Carter, J. Oliver, and T.F. Smith, 2012:
Socio-economic trends and climate change adaptation: the case of South East
Queensland. Australasian Journal of Environmental Management, 19(1), 35-
50.
Rolfe, J., 2009: Climate change and security: the defence component. In: Climate
Change and Security. Planning for the Future [Boston, J., P. Nel, and M. Righarts,
(eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington,
New Zealand, pp. 93-100.
Roman, C.E., A.H. Lynch, and D. Dominey-Howes, 2010: Uncovering the essence of
the climate change adaptation problem a case study of the tourism sector at
Alpine Shire, Victoria, Australia. Tourism Planning & Development, 7(3), 237-
252.
Ross, H., C. Grant, C.J. Robinson, A. Izurieta, D. Smyth, and P. Rist, 2009: Co-management
and Indigenous protected areas in Australia: achievements and ways forward.
Australasian Journal of Environmental Management, 16(4), 242-252.
Rosselló-Nadal, J., A. Riera-Font, and V. Cárdenas, 2011: The impact of weather
variability on British outbound flows. Climatic Change, 105(1), 281-292.
1434
Chapter 25 Australasia
25
Roura-Pascual, N., C. Hui, T. Ikeda, G. Leday, D.M. Richardson, S. Carpintero, X.
Espadaler, C. Gómez, B. Guénard, S. Hartley, P. Krushelnycky, P.J. Lester, M.A.
McGeoch, S.B. Menke, J.S. Pedersen, J.P.W. Pitt, J. Reyes, N.J. Sanders, A.V. Suarez,
Y. Touyama, D. Ward, P.S. Ward, and S.P. Worner, 2011: Relative roles of climatic
suitability and anthropogenic influence in determining the pattern of spread
in a global invader. Proceedings of the National Academy of Sciences of the
United States of America, 108(1), 220-225.
Rouse, H.L. and N. Norton, 2010: Managing scientific uncertainty for resource
management planning in New Zealand. Australasian Journal of Environmental
Management, 17(2), 66-76.
Rouse, H.L. and P. Blackett, 2011: Coastal Adaptation to Climate Change. Engaging
Communities: Making It Work. NIWA Project Report CACC125, National Institute
of Water and Atmospheric Research (NIWA), Christchurch, New Zealand, 75 pp.
Ruddell, A.R., 1995: Recent Glacier and Climate Change in the New Zealand Alps.
PhD thesis, School of Earth Sciences, University of Melbourne, Melbourne, VIC,
Australia, 402 pp.
Russell, B.D., J.A.I. Thompson, L.J. Falkenberg, and S.D. Connell, 2009: Synergistic
effects of climate change and local stressors: CO
2
and nutrient-driven change
in subtidal rocky habitats. Global Change Biology, 15(9), 2153-2162.
Russell, S.L., A. Greenaway, F. Carswell, and S. Weaver, 2013: Moving beyond
“mitigation and adaptation”: examining climate change responses in New
Zealand. Local Environment: The International Journal of Justice and
Sustainability, doi:10.1080/13549839.2013.792047.
Russell-Smith, J., B.P. Murphy, C.P. Meyer, G.D. Cook, S. Maier, A.C. Edwards, J. Schatz,
and P. Brocklehurst, 2009: Improving estimates of savanna burning emissions
for greenhouse accounting in northern Australia: limitations, challenges,
applications. International Journal of Wildland Fire, 18(1), 1-18.
Rutledge, D.T., R.J. Sinclair, A. Tait, J. Poot, M. Dresser, S. Greenhalgh, and M. Cameron,
2011: Triggers and Thresholds of Land-Use Change in Relation to Climate Change
and Other Key Trends: A Review and Assessment of Potential Implications for
New Zealand. Report prepared for the New Zealand Ministry of Agriculture and
Forestry (MAF) by Landcare Research New Zealand Ltd., MAF, Wellington, New
Zealand, 114 pp.
Sadras, V.O. and P.R. Petrie, 2011: Climate shifts in south-eastern Australia: early
maturity of Chardonnay, Shiraz and Cabernet Sauvignon is associated with
early onset rather than faster ripening. Australian Journal of Grape and Wine
Research, 17(2), 199-205.
Salinger, M.J., J.A. Renwick, and A.B. Mullan, 2001: Interdecadal Pacific Oscillation and
South Pacific climate. International Journal of Climatology, 21(14), 1705-1721.
Salinger, M.J., G. Griffiths, and A. Gosai, 2005: Extreme pressure differences at 0900
NZST and winds across New Zealand. International Journal of Climatology,
25(9), 1203-1222.
Sanders, D., J. Laing, and M. Houghton, 2008: Impact of Bushfires on Tourism and
Visitation in Alpine National Parks. Sustainable Tourism Cooperative Research
Centre (CRC), Gold Coast, QLD, Australia, 42 pp.
Sato, Y., D.G. Bourne, and B.L. Willis, 2009: Dynamics of seasonal outbreaks of black
band disease in an assemblage of Montipora species at Pelorus Island (Great
Barrier Reef, Australia). Proceedings of the Royal Society B, 276(1668), 2795-
2803.
Sattler, P. and M. Taylor, 2008: Building Nature’s Safety Net 2008: Progress on the
Directions for the National Reserve System. WWF-Australia Report, World Wide
Fund for Nature-Australia (WWF-Australia), Sydney, NSW, Australia, 64 pp.
Saunders, C., W. Kaye-Blake, L. Marshall, S. Greenhalgh, and M. de Aragao Pereira,
2009: Impacts of a United States’ biofuel policy on New Zealand’s agricultural
sector. Energy Policy, 37(9), 3448-3454.
Saunders, C., W. Kaye-Blake, and J. Turner, 2010: Modelling Climate Change Impacts
on Agriculture and Forestry with the Extended LTEM (Lincoln Trade and
Environment Model). AERU Research Report No. 316, Prepared for the Ministry
of Agriculture and Forestry by the Agribusiness and Economics Research Unit
(AERU), Lincoln University, Lincoln, New Zealand, 53 pp.
Saunders, D.A., P. Mawson, and R. Dawson, 2011: The impact of two extreme
weather events and other causes of death on Carnaby’s black cockatoo: a
promise of things to come for a threatened species? Pacific Conservation
Biology, 17, 141-148.
SCCCWEA, 2009: Managing our Coastal Zone in a Changing Climate: The Time to
Act is Now. House of Representatives Standing Committee on Climate Change,
Water, Environment and the Arts (SCCCWEA) report on the inquiry into climate
change and environmental impacts on coastal communities, Parliament of
Australia, Canberra, ACT, Australia, 368 pp.
Scealy, B., D. Newth, D. Gunasekera, and J. Finnigan, 2012: Effects of variation in the
grains sector response to climate change: an integrated assessment. Economic
Papers, 31(3), 327-336.
Schiel, D.R., 2011: Biogeographic patterns and long-term changes on New Zealand
coastal reefs: non-trophic cascades from diffuse and local impacts. Journal of
Experimental Marine Biology and Ecology, 400(1-2), 33-51.
Schiff, A. and S. Becken, 2011: Demand elasticities for tourism in New Zealand.
Tourism Management, 32(3), 564-575.
Schofield, N., 2011: Climate change and its impacts current understanding, future
directions. In: Basin Futures: Water Reform in the Murray-Darling Basin
[Grafton, Q. and D. Connell (eds.)]. The Australian National University (ANU)
E-Press, Canberra, ACT, Australia, pp. 81-100.
Schrobback, P., D. Adamson, and J. Quiggin, 2011: Turning water into carbon: carbon
sequestration and water flow in the Murray-Darling Basin. Environmental and
Resource Economics, 49(1), 23-45.
Schumann, N., N.J. Gales, R.G. Harcourt, and J.P.Y. Arnould, 2013: Impacts of climate
change on Australian marine mammals. Australian Journal of Zoology, 61, 146-
159.
Schuster, S., 2013: Natural hazards and insurance. In: Climate Adaptation Futures
[Palutikof, J., S. Boulter, A. Ash, M. Stafford-Smith, M. Parry, M. Waschka, and D.
Guitart (eds.)]. John Wiley & Sons, Brisbane, QLD, Australia, pp. 133-140.
Scott, D., S. Gössling, and C.M. Hall, 2012: International tourism and climate change.
Wiley Interdisciplinary Reviews: Climate Change, 3(3), 213-232.
SCRGSP, 2011: Overcoming Indigenous Disadvantage: Key Indicators 2011. Steering
Committee for the Review of Government Service Provision (SCRGSP),
Productivity Commission, Canberra, ACT, Australia, 830 pp.
Scutella, R., R. Wilkins, and W. Kostenko, 2009: Estimates of Poverty and Social
Exclusion in Australia: A Multidimensional Approach. MIAESR Working Paper
No. 26/09, Melbourne Institute of Applied Economic and Social Research
(MIAESR), Melbourne University, Melbourne, VIC, Australia, 68 pp.
Serrao-Neumann, S., F. Crick, B. Harman, M. Sano, O. Sahin, R. Staden, G. Schuch, S.
Baum, and D. Low Choy, 2014: Improving cross-sectoral climate change
adaptation for coastal settlements: insights from South East Queensland,
Australia. Regional Environmental Change, 14(2), 489-500.
SGS, 2012: Tasmanian Coastal Adaptation Pathways Project. Lauderdale Recommended
Actions. Report prepared on behalf of Tasmanian Coastal Adaptation Pathways
Project (TCAP) and Clarence City Council by SGS Economics & Planning, Pty
Ltd., Local Government Association of Tasmania, Hobart, TAS, Australia, 46 pp.
Shakesby, R.A., P.J. Wallbrink, S.H. Doerr, P.M. English, C.J. Chafer, G.S. Humphreys,
W.H. Blake, and K.M. Tomkins, 2007: Distinctiveness of wildfire effects on soil
erosion in south-east Australian eucalypt forests assessed in a global context.
Forest Ecology and Management, 238(1-3), 347-364.
ShapeNZ, 2009: New Zealanders’ Attitudes to Climate Change. National Climate
Change Survey of 2,851 New Zealanders conducted by ShapeNZ for the New
Zealand Business Council for Sustainable Development, Auckland, New
Zealand, 34 pp.
Sharma, V., S.v.d. Graaff, B. Loechel, and D. Franks, 2013: Extractive Resource
Development in a Changing Climate: Learning the Lessons from Extreme
Weather Events in Queensland, Australia. Final Report prepared by the Centre
for Social Responsibility in Mining, Sustainable Minerals Institute, and The
University of Queensland, National Climate Change Adaptation Research
Facility (NCCARF), Griffith University, Gold Coast Campus, Southport, QLD,
Australia, 110 pp.
Shearer, H., 2011: Using Geographic Information Systems to explore the determinants
of household water consumption and response to the Queensland Government
demand-side policy measures imposed during the drought of 2006-2008. In:
5
th
State of Australian Cities National Conference, 29 November-2 December
2011, Melbourne, VIC, Australia [Whitzman, C. and R. Fincher (eds.)]. Melbourne
Organizing Committee and State of Australian Cities Conference Research
Network (SOACRN), Melbourne, VIC, Australia, soac.fbe.unsw.edu.au/2011/
papers/SOAC2011_0140_Final.pdf.
Sheppard, C.S., 2012: Potential spread of recently naturalised plants in New Zealand
under climate change. Climatic Change, 117(4), 919-931.
Shoo, L.P., C. Storlie, J. Vanderwal, J. Little, and S.E. Williams, 2011: Targeted protection
and restoration to conserve tropical biodiversity in a warming world. Global
Change Biology, 17(1), 186-193.
Simioni, G., P. Ritson, M.U.F. Kirschbaum, J. McGrath, I. Dumbrell, and B. Copeland,
2009: The carbon budget of Pinus radiata plantations in south-western Australia
under four climate change scenarios. Tree Physiology, 29(9), 1081-1093.
1435
Australasia Chapter 25
25
Sinclair, E., 2008: The Changing Climate of New Zealand’s Security: Risk and
Resilience in a Climate Affected Security Environment. Working Paper 08/11,
Institute of Policy Studies, Victoria University, Wellington, New Zealand, 68 pp.
Singh, S., S. Davey, and M. Cole, 2010: Implications of climate change for forests,
vegetation and carbon in Australia. New Zealand Journal of Forestry Science,
40, 141-152.
Skinner, R., 2010: Adaptation to Climate Change in Melbourne: Changing the
Fundamental Planning Assumptions. Presented at the International Adaptation
Forum on Climate Change Impacts on Water Supply, Washington DC, January,
2010, Melbourne Water, Melbourne, VIC, Australia, 24 pp.
Smale, D.A. and T. Wernberg, 2013: Extreme climatic event drives range contraction
of a habitat-forming species. Proceedings of the Royal Society B, 280(1754),
20122829, doi:10.1098/rspb.2012.2829.
Smales, I., 2006: Wind Farm Collision Risk for Birds. Cumulative Risks for Threatened
and Migratory Species. Report prepared by Biosis Research Pty Ltd. for the
Australian Government Department of Environment and Heritage, Canberra,
ACT, Australia, 237 pp.
Smart, A. and A. Aspinall, 2009: Water and the Electricity Generation Industry. Waterlines
Report No. 18, National Water Commission, Canberra, ACT, Australia, 101 pp.
Smart, G.M. and A.I. McKerchar, 2010: More flood disasters in New Zealand. Journal
of Hydrology (New Zealand), 49(2), 69-78.
Smart, R.E., 2010: A lump of coal, a bunch of grapes. Journal of Wine Research, 21(2),
107-111.
Smith, H.G., G.J. Sheridan, P.N.J. Lane, P. Nyman, and S. Haydon, 2011: Wildfire effects
on water quality in forest catchments: a review with implications for water
supply. Journal of Hydrology, 396(1-2), 170-192.
Smith, I.N. and B. Timbal, 2010: Links between tropical indices and southern
Australian rainfall. International Journal of Climatology, 32(1), 33-40.
Smith, T.F., C. Brooke, T.G. Measham, B. Preston, R. Gorddard, G. Withycombe, B.
Beveridge, and C. Morrison, 2008: Case Studies of Adaptive Capacity: Systems
Approach to Regional Climate Change Adaptation Strategies. Prepared by the
Commonwealth Scientific and Industrial Research Organization (CSIRO) for the
Sydney Coastal Councils Group Inc., Sydney, NSW, Australia, 110 pp.
Smith, T., T. Lynam, B. Preston, J. Matthews, R.W. Carter, D. Thomsen, J. Carter, A. Roiko,
R. Simpson, P. Waterman, M. Bussey, N. Keys, and C. Stephenson, 2010: Towards
enhancing adaptive capacity for climate change response in South East
Queensland. The Australasian Journal of Disaster and Trauma Studies, 2010-1,
www.massey.ac.nz/~trauma/issues/2010-1/tsmith.htm.
Smith, T., D. Low Choy, D.C. Thomsen, S. Neumann, F. Crick, M. Sano, R. Richards, B.
Harman, S. Baum, S. Myers, V. Sharma, M. Bussey, J. Matthews, A. Roiko, and
R.W. Carter, 2013: Adapting Australian coastal regions to climate change: a
case study of South East Queensland. In: Climate Change and the Coastal Zone:
Building Resilient Communities [Glavovic, B., R.C. Kay, M. Kelly, and A. Travers
(eds.)]. CRC Press, London, UK (in press), www.crcpress.com/product/isbn/
9780415464871.
Smith, W., C. Davies-Colley, A. Mackay, and G. Bankoff, 2011: Social impact of the
2004 Manawatu floods and the ‘hollowing out’ of rural New Zealand. Disasters,
35(3), 540-553.
SNZ, 2010a: National Ethnic Population Projections: 2006 (Base) 2026 Update.
Statistics New Zealand (SNZ), Wellington, New Zealand, 31 pp.
SNZ, 2010b: Are New Zealanders Living Closer to the Coast? Internal Migration.
Statistics New Zealand (SNZ), Wellington, New Zealand, 4 pp.
SNZ, 2010c: Subnational Population Projections: 2006 (Base) 2031 Update. Statistics
New Zealand (SNZ), Wellington, New Zealand, 15 pp.
SNZ, 2011: National Accounts: Long-Term Data Series, Table E.1.2: Real Gross
Domestic Product. Excel data spreadsheet, Statistics New Zealand (SNZ),
Wellington, New Zealand, www.stats.govt.nz.
SNZ, 2012a: National Population Projections: 2011 (Base) 2061. Statistics New
Zealand (SNZ), Wellington, New Zealand, 28 pp.
SNZ, 2012b: Global New Zealand International Trade, Investment, and Travel Profile:
Year ended June 2012. Ministry of Foreign Affairs and Trade and Statistics New
Zealand (SNZ), Wellington, New Zealand, 196 pp.
SNZ, 2013: New Zealand’s Seafood Industry Infographic. Statistics New Zealand
(SNZ), Wellington, New Zealand, 1 pp., www.stats.govt.nz/tools_and_services/
newsletters/price-index-news/apr-13-infographic-seafood.aspx.
SoE, 2011: State of the Environment 2011. Independent Report to the Australian
Government Minister for Sustainability, Environment, Water, Population and
Communities, Australian State of the Environment (SoE) Committee, Canberra,
ACT, Australia, 932 pp.
Soste, L., 2010: Victorian Climate Change Adaptation Program (VCCAP): Final Report
Scenario Theme. Department of Primary Industries, State Government of
Victoria, Tatura, VIC, Australia, 23 pp.
Spickett, J., H. Brown, and D. Katscherian, 2008: Health Impacts of Climate Change:
Adaptation Strategies for Western Australia. Produced by Environmental Health
Directorate, Western Australia Department of Health, Perth, WA, Australia, 74 pp.
Spillman, C.M., 2011: Operational real-time seasonal forecasts for coral reef
management. Journal of Operational Oceanography, 4(1), 13-22.
Srinivasan, M.S., J. Schmidt, S. Poyck, and E. Hreinsson, 2011: Irrigation reliability
under climate change scenarios: a modeling investigation in a river-based
irrigation scheme in New Zealand. Journal of the American Water Resources
Association, 47(6), 1261-1274.
Stafford-Smith, M., 2013: Scenarios for picturing a future adapted to +4°C. In:
Climate Adaptation Futures [Palutikof, J., S.L. Boulter, A.J. Ash, M. Stafford-Smith,
M. Parry, M. Waschka, and D. Guitart (eds.)]. John Wiley & Sons, Brisbane, QLD,
Australia, pp. 119-125.
Stafford-Smith, M., L. Horrocks, A. Harvey, and C. Hamilton, 2011: Rethinking
adaptation for a 4°C world. Philosophical Transactions of the Royal Society A,
369(1934), 196-216.
Stamatov, V. and A. Stamatov, 2010: Long-term impact of water desalination plants
on the energy and carbon dioxide balance of Victoria, Australia: a case study
from Wonthaggi. Water and Environment Journal, 24(3), 208-214.
Standards Australia, 2013: Climate Change Adaptation for Settlements and
Infrastructure – A Risk Based Approach. AS 5334-2013, Standards Australia,
Canberra, ACT, Australia, 47 pp.
Stark, C., K. Penney, and A. Feng, 2012: 2012 Australian Energy Update. Bureau of
Resources and Energy Economics (BREE), Canberra, ACT, Australia, 26 pp.
Steffen, W., A.A. Burbidge, L. Hughes, R. Kitching, D. Lindenmayer, W. Musgrave, M.
Stafford-Smith, and P.A. Werner, 2009: Australias Biodiversity and Climate
Change: A Strategic Assessment of the Vulnerability of Australia’s Biodiversity
to Climate Change. Report commissioned by the Australian Government for
the Natural Resource Management Ministerial Council, Prepared by the
Biodiversity and Climate Change Expert Advisory Group, Commonwealth
Scientific and Industrial Research Organization (CSIRO) Publishing, Collingwood,
VIC, Australia, 290 pp.
Steffen, W., J. Sims, J. Walcott, and G. Laughlin, 2011: Australian agriculture: coping
with dangerous climate change. Regional Environmental Change, 11(Suppl. 1),
205-214.
Stehlik, D., G. Lawrence, and I. Gray, 2000: Gender and drought: experiences of
Australian women in the drought of the 1990s. Disaster, 24(1), 38-53.
Stephenson, J., B. Barton, G. Carrington, D. Gnoth, R. Lawson, and P. Thorsnes, 2010:
Energy cultures: a framework for understanding energy behaviours. Energy
Policy, 38(10), 6120-6129.
Stewart, H.T.L., D.H. Race, and A.L. Curtis, 2011: New forests in changing landscapes
in south-east Australia. International Forestry Review, 13(1), 67-79.
Stewart, M.G. and X. Wang, 2011: Risk Assessment of Climate Adaptation Strategies
for Extreme Wind Events in Queensland. Commonwealth Scientific and
Industrial Research Organization (CSIRO), Canberra, ACT, Australia, 85 pp.
Stewart, M.G., X. Wang, and M.N. Nguyen, 2012: Climate change adaptation for
corrosion control of concrete infrastructure. Structural Safety, 35(0), 29-39.
Stokes, C. and A. Ash, 2007: Impacts of climate change on marginal tropical animal
production systems. In: Agroecosystems in a Changing Climate [Newton, P.C.D.,
R.A. Carran, G.R. Edwards, and P.A. Niklaus (eds.)]. CRC Press, New York, NY,
USA, pp. 323-328.
Stork, N.E., J. Balston, G.D. Farquhar, P.J. Franks, J.A.M. Holtum, and M.J. Liddell, 2007:
Tropical rainforest canopies and climate change. Austral Ecology, 32(1), 105-
112.
Strengers, Y., 2008: Comfort expectations: the impact of demand-management
strategies in Australia. Building Research and Information, 36(4), 381-391.
Strengers, Y., and C. Maller, 2011: Integrating health, housing and energy policies:
social practices of cooling. Building Research and Information, 39(Compendex),
154-168.
Stroombergen, A., 2010: The International Effects of Climate Change on Agricultural
Commodity Prices, and the Wider Effects on New Zealand. Motu Working Paper
10-14, Motu Economic and Public Policy Research, Wellington, New Zealand,
27 pp.
Stroombergen, A., A. Tait, J. Renwick, and K. Patterson, 2006: The relationship
between New Zealand’s climate, energy, and the economy in 2025. Kōtuitui:
New Zealand Journal of Social Sciences Online, 1, 139-160.
1436
Chapter 25 Australasia
25
Suncorp, 2013: Risky Business. Insurance and Natural Disaster Risk Management.
Suncorp Personal Insurance Public Policy. Suncorp Group, Brisbane, QLD,
Australia, 19 pp.
Sutherst, R.W., B.S. Collyer, and T. Yonow, 2000: The vulnerability of Australian
horticulture to the Queensland fruit fly, Bactrocera (Dacus) tryoni, under climate
change. Australian Journal of Agricultural Research, 51(4), 467-480.
Sutton, P.J.H., M. Bowen, and D. Roemmich, 2005: Decadal temperature changes in
the Tasman sea. New Zealand Journal of Marine and Freshwater Research,
39(6), 1321-1329.
Syed, A., 2012: Australian Energy Projections to 2049-50. Bureau of Resources and
Energy Economics (BREE), Canberra, ACT, Australia, 60 pp.
Tait, A., 2008: Future projections of growing degree days and frost in New Zealand
and some implications for grape growing. Weather and Climate, 28, 17-36.
Tait, A., T. Baisden, D. Wratt, B. Mullan, and A. Stroombergen, 2008a: An initial
assessment of the potential effects of climate change on New Zealand
agriculture. New Zealand Science Review, 65(3), 50-56.
Tait, A., J. Sturman, B. Mullan, D. King, G. Griffiths, P. Newsome, G. Harmsworth, I.
Nicholas, L. Gea, N. Porter, and J. Reid, 2008b: Use of Climate, Soil, and Tree
Species Information for Identifying Land Use Options Across the Gisborne
District. NIWA Client Report: AKL-2007-038, Prepared for Ngati Porou Whanui
Forests Ltd. by National Institute of Water and Atmospheric Research (NIWA),
Landcare Research Ltd., Crop & Food Research, and SCION / ENSIS Research
Ltd., NIWA, Wellington, New Zealand, 104 pp.
Taptiklis, N., 2011: Climate Resilient Water Management in Wellington, New Zealand.
NZCCRI 2011 Report 09, New Zealand Climate Change Research Institute
(NZCCRI), Victoria University of Wellington, Wellington, New Zealand, 62 pp.
Taylor, B.M., B.P. Harman, S. Heyenga, and R.R.J. McAllister, 2012: Property developers
and urban adaptation: conceptual and empirical perspectives on governance.
Urban Policy and Research, 30(1), 5-24.
Taylor, M.A.P. and M. Philp, 2010: Adapting to climate change implications for
transport infrastructure, transport systems and travel behaviour. Road &
Transport Research, 19(4), 66-79.
Taylor, R., D. Rutledge, and H. vanRoon, 2011: Building capacity in urban sustainability
assessment through use of a scenarios game. Journal of Education for Sustainable
Development, 5(1), 75-87.
Taylor, S., L. Kumar, and N. Reid, 2012: Impacts of climate change and land-use on
the potential distribution of an invasive weed: a case study of Lantana camara
in Australia. Weed Research, 52(5), 391-401.
Te Aho, L., 2007: Contemporary issues in Māori law and society. Waikato Law
Review, 15, 138-158.
Teixeira, E.I., H.E. Brown, A. Fletcher, G. Hernandez-Ramirez, A. Soltani, S. Viljanen-
Rollinson, A. Horrocks, and P. Johnstone, 2012: Broadacre cropping: adapting
broad acre cropping to climate change. In: Enhanced Climate Change Impact
and Adaptation Evaluation: A Comprehensive Analysis of New Zealand’s Land-
Based Primary Sectors [Clark, A. and R. Nottage (eds.)]. Ministry for Primary
Industries, Wellington, New Zealand, pp. 189-236.
Teng, J., F. Chiew, J. Vaze, S. Marvanek, and D.G.C. Kirono, 2012: Estimation of climate
change impact on mean annual runoff across continental Australia using
Budyko and Fu equations and comparison with hydrological model simulations.
Journal of Hydrometeorology, 13, 122-139.
TGA, 2009: The Geneva Reports – Risk and Insurance Reseach No. 2. The Insurance
Industry and Climate Change: Contribution to the Global Debate. The Geneva
Association (TGA), Geneva, Switzerland, 138 pp.
Thatcher, M.J., 2007: Modelling changes to electricity demand load duration
curves as a consequence of predicted climate change for Australia. Energy,
32(Compendex), 1647-1659.
Thompson, D. and J. Wallace, 2000: Annular modes in the extratropical circulation.
Part I: month-to-month variability. Journal of Climate, 13(5), 1000-1016.
Thomas, D.T., J. Sanderman, S.J. Eady, D.G. Masters, and P. Sanford, 2012: Whole farm
net greenhouse gas abatement from establishing Kikuyu-based perennial
pastures in south-western Australia. Animals, 2(3), 316-330.
Thompson, P.A., M.E. Baird, T. Ingleton, and M.A. Doblin, 2009: Long-term changes
in temperate Australian coastal waters: implications for phytoplankton. Marine
Ecology Progress Series, 394, 1-19.
Thomson, L.J., S. Macfadyen, and A.A. Hoffmann, 2010: Predicting the effects of
climate change on natural enemies of agricultural pests. Biological Control,
52(3), 296-306.
Thorburn, P.J., M.J. Robertson, B.E. Clothier, V.O. Snow, E. Charmley, J. Sanderman, E.
Teixeira, R.A. Dynes, A. Hall, H. Brown, S.M. Howden, and M. Battaglia, 2012:
Australia and New Zealand perspectives on climate change and agriculture. In:
Agricultures Contributions to Climate Change Solutions: Mitigation and
Adaptation at Global and Regional Scales. Handbook of Climate Change and
Agroecosystems, Vol 2 [Rosenzweig, C. and D. Hillel (eds.)]. American Society
of Agronomy and Imperial College Press, New York, NY, USA, pp. 107-141.
Thresher, R.E., J.A. Koslow, A.K. Morison, and D.C. Smith, 2007: Depth-mediated
reversal of the effects of climate change on long-term growth rates of exploited
marine fish. Proceedings of the National Academy of Sciences of the United
States of America, 104(18), 7461-7465.
Timbal, B. and D.A. Jones, 2008: Future projections of winter rainfall in southeast
Australia using a statistical downscaling technique. Climatic Change, 86(1-2),
165-187.
Timbal, B., J.M. Arblaster, and S. Power, 2006: Attribution of the late-twentieth-century
rainfall decline in southwest Australia. Journal of Climate, 19(10), 2046-2062.
Timbal, B., J. Arblaster, K. Braganza, E. Fernandez, H. Hendon, B. Murphy, M. Raupach,
C. Rakich, I. Smith, K. Whan, and M. Wheeler, 2010a: Understanding the
Anthropogenic Nature of the Observed Rainfall Decline Across South Eastern
Australia. CAWCR Technical Report No. 26, Centre for Australian Weather and
Climate Research (CAWCR), partnership between the Bureau of Meteorology
and CSIRO, CAWCR, Melbourne, VIC, Australia, 180 pp.
Timbal, B., R. Kounkou, and G.A. Mills, 2010b: Changes in the risk of cool-season
tornadoes over southern Australia due to model projections of anthropogenic
warming. Journal of Climate, 23(9), 2440-2449.
Todd, M., W. Zhang, and S. Kerr, 2009: Competition for land between biofuels,
pastoral agriculture. In: Bioenergy Options for New Zealand: Analysis of Large-
Scale Bioenergy from Forestry [Hall, P. and M. Jack (eds.)]. Scion, Rotorua, New
Zealand, pp. 122-140.
Tolhurst, K., 2009: Report of the Physical Nature of the Victorian Fires occurring on
7 February 2009. Report for the Counsel Assisting the Royal Commission,
Victorian Bushfire Royal Commission, Melbourne, VIC, Australia, 18 pp.
Tong, S.L., C.Z. Ren, and N. Becker, 2010a: Excess deaths during the 2004 heatwave
in Brisbane, Australia. International Journal of Biometeorology, 54(4), 393-400.
Tong, S.L., X.Y. Wang, and A.G. Barnett, 2010b: Assessment of heat-related health
impacts in Brisbane, Australia: comparison of different heatwave definitions.
PLoS ONE, 5(8), e12155, doi:10.1371/journal.pone.0012155.
Tourism Queensland, 2007: Regional Tourism Crisis Management Plan Template. A
Guide to Preparing a Regional Tourism Crisis Management Plan. Tourism
Queensland, Brisbane, QLD, Australia, 30 pp.
Tourism Queensland, 2010: Climate Futures Industry Tool. Tourism Queensland,
Brisbane, QLD, Australia, 31 pp.
Tourism Victoria, 2010: Crisis Essentials. Crisis Management for Tourism Businesses.
Tourism Victoria, Melbourne, VIC, Australia, 18 pp.
TPK, 2007: A Time for Change in Māori Economic Development. Te Puni Kokiri (TPK),
Wellington, New Zealand, 32 pp.
TRA, 2010: Impact of the Drought on Tourism in the Murray River Region: Summary
of Results. Tourism Research Australia (TRA), Department of Resources, Energy
and Tourism, Canberra, ACT, Australia, 3 pp.
Traill, L.W., K. Perhans, C.E. Lovelock, A. Prohaska, S. McFallan, J.R. Rhodes, and K.A.
Wilson, 2011: Managing for change: wetland transitions under sea-level rise and
outcomes for threatened species. Diversity and Distributions, 17(6), 1225-1233.
Trewin, B. and H. Vermont, 2010: Changes in the frequency of record temperatures
in Australia, 1957-2009. Australian Meteorological and Oceanographic Journal,
60(2), 113-120.
Troccoli, A., K. Muller, P. Coppin, R. Davy, C. Russell, and A.L. Hirsch, 2012: Long-term
wind speed trends over Australia. Journal of Climate, 25, 170-183.
Trotman, R., 2008: Project Twin Streams to 2007. Waitakere City Councils Story.
Waitakere City Council, Waitakere, New Zealand, 56 pp.
Troy, P. (ed.), 2008: Troubled Waters: Confronting the Water Crisis in Australia’s Cities.
The Australian National University (ANU) E-Press, Canberra, ACT, Australia, 217
pp.
Tryhorn, L. and J. Risbey, 2006: On the distribution of heat waves over the Australian
region. Australian Meteorological Magazine, 55(3), 169-182.
TSRA, 2010: Torres Strait Climate Change Strategy 2010-2013. Environmental
Management Program, Torres Strait Regional Authority (TSRA), Darwin, NT,
Australia, 18 pp.
Turton, S., W. Hadwen, R. Wilson, B. Jorgensen, and D. Simmons, 2009: The Impacts
of Climate Change on Australian Tourist Destinations: Developing Adaptation
and Response Strategies. Sustainable Tourism Cooperative Research Centre,
Gold Coast, Qld, Australia, 77 pp.
1437
Australasia Chapter 25
25
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), 429-448.
UN DESA Population Division, 2012: World Urbanization Prospects: The 2011
Revision. United Nations Department of Economic and Social Affairs Population
Division, United Nations (UN) Secretariat, New York, NY, USA, 302 pp.
Valenzuela, E. and K. Anderson, 2011: Climate change and food security to 2030: a
global economy-wide perspective. Economia Agraria y Recursos Naturales,
11(1), 29-58.
van den Honert, R.C. and J. McAneney, 2011: The 2011 Brisbane Floods causes,
impacts and implications. Water, 3, 1149-1173.
van der Wal, J., H.T. Murphy, A.S. Kutt, G.C. Perkins, B.L. Bateman, J.J. Perry, and
A.R. Reside, 2013: Focus on poleward shifts in species’ distribution
underestimates the fingerprint of climate change. Nature Climate Change, 3,
239-243.
van Dijk, A., R. Evans, P. Hairsine, S. Khan, R. Nathan, Z. Paydar, N. Viney, and L. Zhang,
2006: Risks to Shared Water Resources of the Murray-Darling Basin. MDBC
Publication 22/06, Prepared for the Murray Darling Basin Commission (MDBC)
by CSIRO Land and Water as part of the Water for a Healthy Country National
Flagship Program, including contributions from Rory Nathan of Sinclair Knight
Merz and Ray Evans of Salient Solutions Inc., MDBC, Canberra, ACT, Australia,
49 pp.
Vaneckova, P., P.J. Beggs, R.J. de Dear, and K.W.J. McCracken, 2008: Effect of
temperature on mortality during the six warmer months in Sydney, Australia,
between 1993 and 2004. Environmental Research, 108(3), 361-369.
VBRC, 2010: 2009 Victorian Bushfires Royal Commission Final Report Summary.
2009 Victorian Bushfires Royal Commission (VBRC), Melbourne, VIC, Australia,
42 pp.
Veland, S., R. Howitt, and D. Dominey-Howes, 2010: Invisible institutions in emergencies:
evacuating the remote Indigenous community of Warruwi, Northern Territory
Australia, from Cyclone Monica. Environmental Hazards: Human and Policy
Dimensions, 9(2), 197-214.
Verdon, D.C., A.M. Wyatt, A.S. Kiem, and S.W. Franks, 2004: Multidecadal variability
of rainfall and streamflow: Eastern Australia. Water Resources Research,
40(10), W10201, doi:10.1029/2004WR003234.
Verdon-Kidd, D.C., A.S. Kiem, and E.K. Austin, 2012: Decision Making Under
Uncertainty – Bridging the Gap Between End User Needs and Climate Science
Capability. National Climate Change Adaptation Research Facility (NCCARF),
Griffith University, Gold Coast Campus, Southport, QLD, Australia, 116 pp.
Veron, J.E.N., O. Hoegh-Guldberg, T.M. Lenton, J.M. Lough, D.O. Obura, P. Pearce-
Kelly, C.R.C. Sheppard, M. Spalding, M.G. Stafford-Smith, and A.D. Rogers, 2009:
The coral reef crisis: the critical importance of <350 ppm CO
2
. Marine Pollution
Bulletin, 58(10), 1428-1436.
Verschuuren, J. and J. McDonald, 2012: Towards a legal framework for coastal
adaptation: assessing the first steps in Europe and Australia. Transnational
Environmental Law, 1(2), 355-379.
Victorian Government, 2009a: January 2009 Heatwave in Victoria: An Assessment
of Health Impacts. Department of Human Services, Melbourne, VIC, Australia,
16 pp.
Victorian Government, 2009b: Heatwave Plan for Victoria 2009-2010: Protecting
Health and Reducing Harm from Heatwaves. Department of Health, Melbourne,
VIC, Australia, 42 pp.
Walker, S.J., T.A. Schlacher, and L.M.C. Thompson, 2008: Habitat modification in a
dynamic environment: the influence of a small artificial groyne on macrofaunal
assemblages of a sandy beach. Estuarine, Coastal and Shelf Science, 79(1), 24-
34.
Wan, S., R.J. Norby, J. Ledford, and J.F. Weltzin, 2007: Responses of soil respiration to
elevated CO
2
, air warming, and changing soil water availability in an old-field
grassland. Global Change Biology, 13(11), 2411-2424.
Wang, X., D. Chen, and Z.G. Ren, 2010a: Assessment of climate change impact on
residential building heating and cooling energy requirement in Australia.
Building and Environment, 45(7), 1663-1682.
Wang, X., M. Stafford-Smith, R. McAllister, A. Leitch, S. McFallan, and S. Meharg,
2010b: Coastal Inundation Under Climate Change: A Case Study in South East
Queensland. Report prepared for the South East Queensland Climate Adaptation
Research Initiative, CSIRO Climate Adaptation Flagship Working Paper No. 6,
Commonwealth Scientific and Industrial Research Organization (CSIRO), Brisbane,
QLD, Australia, 28 pp.
Wang, X. and R. McAllister, 2011: Adapting to heatwaves and coastal flooding. In:
Climate Change: Science and Solutions for Australia [Cleugh, H., M. Stafford-
Smith, M. Battaglia, and P. Graham (eds.)]. Commonwealth Scientific and
Industrial Research Organization (CSIRO) Publishing, Collingwood, VIC,
Australia, pp. 73-84.
Wang, X., M. Stewart, M. Nguyen, and M. Syme, 2012: Implications of climate change
to the durability of concrete structures an investigation by simulations. Concrete
in Australia, 36(4), 27-35.
Watanabe, T., S. Bowatte, and P.C.D. Newton, 2013: A reduced fraction of plant N
derived from atmospheric N (%Ndfa) and reduced rhizobial nifH gene numbers
indicate a lower capacity for nitrogen fixation in nodules of white clover
exposed to long-term CO
2
enrichment. Biogeosciences, 10, 8269-8281.
Watson, R.A., G.B. Nowara, S.R. Tracey, E.A. Fulton, C.M. Bulman, G.J. Edgar, N.S.
Barrett, J.M. Lyle, S.D. Frusher, and C.D. Buxton, 2012: Ecosystem model of
Tasmanian waters explores impacts of climate change induced changes in
primary productivity. Ecological Modelling, 264, 115-129.
Watt, M.S., M.U.F. Kirschbaum, T.S.H. Paul, A. Tait, H.G. Pearce, E.G. Brockerhoff, J.R.
Moore, L.S. Bulman, and D.J. Kriticos, 2008: The Effect of Climate Change on
New Zealand’s Planted Forests: Impacts, Risks and Opportunities. Scion Client
Report CC MAF POL 2008-07 (106-1) No. 1, Report produced by Scion for the
Ministry of Agriculture and Forestry, Scion, Rotorua, New Zealand, 149 pp.
Watt, M.S., D.J. Palmer, and L.S. Bulman, 2011a: Predicting the severity of Dothistroma
needle blight on Pinus radiata under climate change. New Zealand Journal of
Forestry Science, 41, 207-215.
Watt, M.S., J.K. Stone, I.A. Hood, and L.K. Manning, 2011b: Using a climatic niche
model to predict the direct and indirect impacts of climate change on the
distribution of Douglas-fir in New Zealand. Global Change Biology, 17(12),
3608-3619.
Watterson, I.G., 2012: Understanding and partioning future climates for Australian
regions from CMIP3 using ocean warming indices. Climatic Change, 111, 903-
922.
WCC, 2010: Wellington City’s 2010 Climate Change Action Plan. Wellington City
Council (WCC), Wellington, New Zealand, 39 pp.
Webb, B. and J.-L. Beh, 2013: Leading Adaptation Practices and Support Strategies
for Australia: An International and Australian Review of Products and Tools.
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University, Gold Coast Campus, Southport, QLD, Australia, 106 pp.
Webb, B., R. McKellar, and R. Kay, 2013: Climate change adaptation in Australia:
experience, challenges and capacity-building. Australian Journal of Emergency
Management, 20(4), 320-337.
Webb, L.B., P.H. Whetton, and E.W.R. Barlow, 2008: Climate change and winegrape
quality in Australia. Climate Research, 36(2), 99-111.
Webb, L.B., P.H. Whetton, J. Bhend, R. Darbyshire, P.R. Briggs, and E.W.R. Barlow,
2012: Earlier wine-grape ripening driven by climatic warming and drying and
management practices. Nature Climate Change, 2, 259-264.
Webb, N.P., C.J. Stokes, and J.C. Scanlan, 2012: Interacting effects of vegetation, soils
and management on the sensitivity of Australian savanna rangelands to climate
change. Climatic Change, 112(3-4), 925-943.
Weber, E.P., A. Memon, and B. Painter, 2011: Science, society, and water resources
in New Zealand: recognizing and overcoming a societal impasse. Journal of
Environmental Policy & Planning, 13(1), 49-69.
Webster, T., J. Morison, N. Abel, E. Clark, L. Rippin, A. Herr, B. Taylor, and P. Stone,
2009: Irrigated agriculture: development opportunities and implications for
northern Australia. In: Northern Australia Land and Water Science Review 2009.
Report to the Northern Territories Land and Water Task Force coordinated by
the Commonwealth Scientific and Industrial Research Organization (CSIRO)
Sustainable Agriculture Flagship, Department of Infrastructure, Transport, Regional
Development and Local Government, Canberra, ACT, Australia, pp. 10-1 to 10-53.
Welbergen, J.A., S.M. Klose, N. Markus, and P. Eby, 2008: Climate change and the
effects of temperature extremes on Australian flying-foxes. Proceedings of the
Royal Society B, 275(1633), 419-425.
Wernberg, T., B.D. Russell, M.S. Thomsen, C.F.D. Gurgel, C.J.A. Bradshaw, E.S.
Poloczanska, and S.D. Connell, 2011: Seaweed communities in retreat from
ocean warming. Current Biology, 21(21), 1828-1832.
Wernberg, T., D.A. Smale, F. Tuya, M.S. Thomsen, T.J. Langlois, T. de Bettignies, S. Bennett,
and C.S. Rousseaux, 2013: An extreme climatic event alters marine ecosystem
structure in a global biodiversity hotspot. Nature Climate Change, 3, 78-82.
Westra, S., 2012: Adapting to Climate Change Revising our Approach to Estimating
Future Floods. University of New South Wales, Sydney, NSW, Australia, 25 pp.
1438
Chapter 25 Australasia
25
Westra, S. and S. Sisson, 2011: Detection of non-stationarity in precipitation extremes
using a max-stable process model. Journal of Hydrology, 406(1-2), 119-128.
Westra, S., J. Evans, R. Mehrotra, and A. Sharma, 2013: A conditional disaggregation
algorithm for generating fine time-scale rainfall data in a warmer climate.
Journal of Hydrology, 479, 86-99.
White, D.A., D.S. Crombie, J. Kinal, M. Battaglia, J.F. McGrath, D.S. Mendham, and S.N.
Walker, 2009: Managing productivity and drought risk in Eucalyptus globulus
plantations in south-western Australia. Forest Ecology and Management, 259,
33-44.
White, N.E., 2009: Local government planning responses to the physical impacts of
climate change in New South Wales, Australia. International Journal of Climate
Change Impacts and Responses, 1(2), 1-16.
White, N.E. and J. Buultjens, 2012: Climate change policy responses of Australia and
New Zealand national governments: implications for sustainable tourism. In:
Tourism, Climate Change and Sustainability [Reddy, M.V. and K. Wilkes (eds.)].
Earthscan Publications, Oxford, UK, pp. 117-133.
Whitehead, P.J., P. Purdon, P.M. Cooke, J. Russell-Smith, and S. Sutton, 2009: The West
Arnhem Land Fire Abatement (WALFA) project: the institutional environment
and its implications. In: Culture, Ecology and Economy of Fire Management in
North Australia Savannas: Rekindling the Wurrk Tradition [Russell-Smith, J., P.J.
Whitehead, and P.M. Cooke (eds.)]. Commonwealth Scientific and Industrial
Research Organization (CSIRO) Publishing, Collingwood, VIC, Australia, pp. 287-
312.
Whittaker, J., J. Handmer, and D. Karoly, 2013: After ‘Black Saturday’: adapting to
bushfires in a changing climate. In: Natural Disasters and Adaptation to Climate
Change [Boulter, S., J. Palutikof, D. Karoly, and D. Guitart (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 115-136.
Wilby, R.L. and R.J. Keenan, 2012: Adapting to flood risk under climate change.
Progress in Physical Geography, 36(3), 348-378.
Wilkinson, S.N., P.J. Wallbrink, G.J. Hancock, W.H. Blake, R.A. Shakesby, and S.H. Doerr,
2009: Fallout radionuclide tracers identify a switch in sediment sources and
transport-limited sediment yield following wildfire in a eucalypt forest.
Geomorphology, 110(3-4), 140-151.
Williams, N.S.G., J.P. Rayner, and K.J. Raynor, 2010: Green roofs for a wide brown
land: opportunities and barriers for rooftop greening in Australia. Urban
Forestry & Urban Greening, 9(3), 245-251.
Williams, R.J., R.A. Bradstock, G.J. Cary, N.J. Enright, A.M. Gill, A.C. Leidloff, C. Lucas,
R.J. Whelan, A.N. Andersen, D.J.M.S. Bowman, P.J. Clarke, G.D. Cook, K.J.
Hennessy, and A. York, 2009: Interactions Between Climate Change, Fire
Regimes and Biodiversity in Australia – A Preliminary Assessment. Report by
CSIRO-led Consortium for the Australian Government Department of Climate
Change (DCC) and Department of the Environment, Water, Heritage and the
Arts (DEWHA), DCC, Canberra, ACT, Australia, 196 pp.
Williams, S.E., L.P. Shoo, J.L. Isaac, A.A. Hoffmann, and G. Langham, 2008: Towards
an integrated framework for assessing the vulnerability of species to climate
change. PLoS Biology, 6(12), 2621-2626.
Willis, T.J., S.J. Handley, F.H. Chang, C.S. Law, D.J. Morrisey, A.B. Mullan, M. Pinkerton,
K.L. Rodgers, P.J.H. Sutton, and A. Tait, 2007: Climate Change and the New
Zealand Marine Environment. NIWA Client Report NEL2007-025, Prepared for
the Department of Conservation by the National Institute of Water and
Atmospheric Research (NIWA), Wellington, New Zealand, 81 pp.
Willsman, A., T. Chinn, J. Hendrikx, and A. Lorrey, 2010: New Zealand Glacier
Monitoring: End of Summer Snowline Survey 2010. NIWA Client Report no.
CHC2010-113, Prepared for the Foundation for Research, Science and Technology
by National Institute of Water and Atmospheric Research (NIWA), Christchurch,
New Zealand, 132 pp.
Wilson, J. and S. Becken, 2011: Perceived deficiencies in the provision of climate and
weather information for tourism: a New Zealand media analysis. New Zealand
Geographer, 67(3), 148-160.
Wilson, N., 2011: End-of-term review of the New Zealand Government’s response
to climate change: a public health perspective. New Zealand Medical Journal,
124(1345), 90-95.
Wilson, R. and S. Turton, 2011a: Climate Change Adaptation Options, Tools and
Vulnerability. Contribution of Work Package 4 to the Forest Vulnerability
Assessment, National Climate Change Adaptation Research Facility (NCCARF),
Griffith University, Gold Coast Campus, Southport, QLD, Australia, 119 pp.
Wilson, R. and S. Turton, 2011b: The impact of climate change on reef-based tourism
in Cairns, Australia – adaptation and response strategies for a highly vulnerable
destination. In: Disappearing Destinations [Jones, A. and M. Phillips (eds.)]. CAB
International, Wallingford, UK, pp. 233-253.
Winterbourn, M.J., S. Cadbury, C. Ilg, and A.M. Milner, 2008: Mayfly production in a
New Zealand glacial stream and the potential effect of climate change.
Hydrobiologia, 603, 211-219.
Wittwer, G. and M. Griffith, 2011: Modelling drought and recovery in the southern
Murray-Darling basin. Australian Journal of Agricultural and Resource Economics,
55(3), 342-359.
Woods, R., A. Tait, B. Mullan, J. Hendrikx, and J. Diettrich, 2008: Projected Climate
and River Flow for the Rangitata Catchment for 2040. NIWA Client Report:
CHC2008-097, Prepared by the National Institute of Water and Atmospheric
Research (NIWA) for Aqualinc Research Ltd., NIWA, Christchurch, New Zealand,
28 pp.
Woodward, A., S. Hales, and N. de Wet, 2001: Climate Change: Potential Effects on
Human Health in New Zealand. Report prepared for the Ministry for the
Environment (MfE) as part of the New Zealand Climate Change Programme,
MfE, Wellington, New Zealand, 21 pp.
Woodward, E., S. Jackson, M. Finn, and P.M. McTaggart, 2012: Utilising Indigenous
seasonal knowledge to understand aquatic resource use and inform water
resource management in northern Australia. Ecological Management &
Restoration, 13(1), 58-64.
Wooldridge, S.A., T.J. Done, C.R. Thomas, I.I. Gordon, P.A. Marshall, and R.N. Jones,
2012: Safeguarding coastal coral communities on the central Great Barrier Reef
(Australia) against climate change: realizable local and global actions Climatic
Change, 112(3-4), 945-961.
Wratt, D., A.B. Mullan, A. Tait, R. Woods, T. Baisden, D. Giltrap, J. Hendy, and A.
Stroombergen, 2008: Costs and Benefits of Climate Change and Adaptation to
Climate Change in New Zealand Agriculture: What Do We Know So Far?
Contract Report by the Ecoclimate Consortium for the Ministry of Agriculture
and Forestry (MAF), MAF, Wellington, New Zealand, 121 pp.
Wyborn, C., 2009: Managing change or changing management: climate change and
human use in Kosciuszko National Park. Australasian Journal of Environmental
Management, 16(4), 208-217.
Yates, C.J., J. Elith, A.M. Latimer, D. Le Maitre, G.F. Midgley, F.M. Schurr, and A.G. West,
2010a: Projecting climate change impacts on species distributions in
megadiverse South African Cape and Southwest Australian Floristic Regions:
opportunities and challenges. Austral Ecology, 35(4), 374-391.
Yates, C.J., A. McNeill, J. Elith, and G.F. Midgley, 2010b: Assessing the impacts of
climate change and land transformation on Banksia in the South West
Australian Floristic Region. Diversity and Distributions, 16(1), 187-201.
Yeates, G. and P.C.D. Newton, 2009: Long-term changes in topsoil nematode
populations in grazed pasture under elevated atmospheric carbon dioxide.
Biology and Fertility of Soils, 45(8), 799-808.
Yuen, E., S. Jovicich, and B. Preston, 2012: Climate change vulnerability assessments
as catalysts for social learning: four case studies in south-eastern Australia.
Mitigation and Adaptation Strategies for Global Change, 18(5), 567-590.
Zammit, C. and R. Woods, 2011: Projected Climate and River Flow for the Ashley
Catchment for 2040 and 2090. NIWA Client Report: CHC2010-160, National
Institute of Water and Atmospheric Research (NIWA), Christchurch, New
Zealand, 46 pp.
Zander, K.K., L. Petheram, and S.T. Garnett, 2013: Stay or leave? Potential climate
change adaptation strategies among Aboriginal people in coastal communities
in northern Australia. Natural Hazards, 67(2), 591-609.
Zeppel, H. and N. Beaumont, 2011: Green tourism futures: climate change responses
by Australian Government tourism agencies. In: CAUTHE 2011: Tourism:
Creating a Brilliant Blend [Gross, M.J. (ed.)]. Proceedings of the 21
st
Annual
Conference:, 8-11 February 2011, University of South Australia, organized by
the Council for Australian University Tourism and Hospitality Education
(CAUTHE), School of Management, University of South Australia, Adelaide, SA,
Australia, pp. 850-865.