FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 25. Australasia Coordinating Lead Authors Andy Reisinger (New Zealand), Roger Kitching (Australia) Lead Authors Francis Chiew (Australia), Lesley Hughes (Australia), Paul Newton (New Zealand), Sandra 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), David Jones (Australia), Sue Jackson (Australia), Darren King (New Zealand), Miko Kirschbaum (New Zealand), Jo Luck (Australia), Jan McDonald (Australia), Kathy McInnes (Australia), Yiheyis Maru (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) Contents Executive Summary 25.1. Introduction and Major Conclusions from Previous Assessments 25.2. Observed and Projected Climate Change 25.3. Socio-Economic Trends Influencing Vulnerability and Adaptive Capacity 25.3.1. Economic, Demographic and Social Trends 25.3.2. Use and Relevance of Socio-Economic Scenarios in Adaptive Capacity/Vulnerability Assessments 25.4. Cross-Sectoral Adaptation: Approaches, Effectiveness, and Constraints 25.4.1. Frameworks, Governance, and Institutional Arrangements 25.4.2. Constraints on Adaptation and Leading Practice Models 25.4.3. Socio-cultural Factors Influencing Impacts of and Adaptation to Climate Change 25.5. Freshwater Resources 25.5.1. Observed Impacts 25.5.2. Projected Impacts 25.5.3. Adaptation 25.6. Natural Ecosystems 25.6.1. Inland Freshwater and Terrestrial Ecosystems 25.6.1.1. Observed Impacts 25.6.1.2. Projected Impacts 25.6.1.3. Adaptation 25.6.2. Coastal and Ocean Ecosystems 25.6.2.1. Observed Impacts Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 25.6.2.2. Projected Impacts 25.6.2.3. Adaptation 25.7. Major Industries 25.7.1. Production Forestry 25.7.1.1. Observed and Projected Impacts 25.7.1.2. Adaptation 25.7.2. Agriculture 25.7.2.1. Projected Impacts and Adaptation – Livestock Systems 25.7.2.2. Projected Impacts and Adaptation – Cropping 25.7.2.3. Integrated Adaptation Perspectives 25.7.3. Mining 25.7.4. Energy Supply, Transmission, and Demand 25.7.5. Tourism 25.7.5.1. Projected Impacts 25.7.5.2. Adaptation 25.8. Human Society 25.8.1. Human Health 25.8.1.1. Observed Impacts 25.8.2.2. Projected Impacts 25.8.3.3. Adaptation 25.8.2. Indigenous Peoples 25.8.2.1. Aboriginal and Torres Strait Islanders 25.8.2.2. New Zealand Māori 25.9. Interactions among Impacts, Adaptation, and Mitigation Responses 25.9.1. Interactions among Local-Level Impacts, Adaptation, and Mitigation Responses 25.9.2. Intra- and Inter-Regional Flow-On Effects between Impacts, Adaptation, and Mitigation 25.10. Synthesis and Regional Key Risks 25.10.1. Economy-wide Impacts and Potential of Mitigation to Reduce Risks 25.10.2. Regional Key Risks as a Function of Mitigation and Adaptation 25.10.3. The Role of Adaptation in Managing Key Risks, and Adaptation limits 25.11. Filling Knowledge Gaps to Improve Management of Climate Risks References Chapter Boxes 25-1. Coastal Adaptation – Planning and Legal Dimensions 25-2. Adaptation through Water Resources Policy and Management in Australia 25-3. Impacts of a Changing Climate in Natural and Managed Ecosystems 25-4. Biosecurity 25-5. Climate Change Vulnerability and Adaptation in Rural Areas 25-6. Climate Change and Fire 25-7. Insurance as Climate Risk Management Tool 25-8. Changes in Flood Risk and Management Responses 25-9. Opportunities, Constraints, and Challenges to Adaptation in Urban Areas 25-10. Land-based Interactions Among Climate, Energy, Water, and Biodiversity Frequently Asked Questions 25.1: How can we adapt to climate change if projected future changes remain uncertain? 25.2: What are the key risks from climate change to Australia and New Zealand? Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 south-western Australia (high confidence) and elsewhere in most of far southern Australia and the north-east 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 13.5, 13.6] 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, Box 25-2, Box 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 north-eastern 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 over 2,000 buildings and led to 173 deaths; and widespread drought in south-east Australia (1997-2009) and many parts of New Zealand (2007-2009; 2012-13) resulted in substantial economic losses (e.g. regional GDP in the southern Murray Darling Basin was below forecast by about 5.7% in 2007/08, and New Zealand lost about NZ$3.6b in direct and off-farm output in 2007-09). [Table 25-1, 25.6.2, 25.8.1, Box 25-5, Box 25-6, Box 25-8] Without adaptation, further changes in climate, atmospheric CO2 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 south-west and far south-east mainland Australia (high confidence) and for rivers originating in the north-east 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, 25.6.2, 25.7.2, 25.7.4, Box 25-1, Box 25-5, Box 25-8] Some sectors in some locations have the potential to benefit from projected changes in climate and increasing atmospheric CO2 (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 Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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, 25.7.2, 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 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, Box 25-1, Box 25-2, Box 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 towards the risks associated with climate change, and different values placed on objects and places at risk. [25.4, 25.10.3, Box 25-1, Table 25-2] Indigenous peoples in both Australia and New Zealand have higher than average exposure to climate change due to 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 engagement 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: o 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] o 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: o 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 [Table 25-1, 25.4.2, Box 25-8, 25.10.3] o constraints on water resources in southern Australia, driven by rising temperatures and reduced coolseason 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] o increased morbidity, mortality and infrastructure damages during heat waves in Australia, resulting from increased frequency and magnitude of extreme high temperatures; vulnerable populations include Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the elderly and those with existing chronic diseases; population increases and ageing trends constrain effectiveness of adaptation responses [25.8.1] o 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, Table 25-1, 25.6.1, 25.7.1, 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: o increasing risks to coastal infrastructure and low-lying ecosystems in Australia and New Zealand from continuing sea level rise, with widespread damages towards the upper end of projected changes; managed retreat is a long-term adaptation strategy for human systems but options for some natural ecosystems are limited due 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 stabilised. [Table 25-1, 25.4.2, Box 25-1, 25.6.1, 25.6.2; WGI 13.5] o significant reduction in agricultural production in the Murray-Darling Basin and far south-eastern and south-western Australia if scenarios of severe drying are realised; 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, Box 25-2, Box 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 (high agreement, limited evidence), but they remain among the least explored issues. [25.7.5, 25.9.1, 25.9.2, Box 25-10] Understanding of future vulnerability of human and mixed human-natural systems to climate change remains limited due to incomplete consideration of socio-economic dimensions (very high confidence). Future vulnerability will depend on factors such as wealth and its distribution across society, patterns of ageing, access to technology and information, labour 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 bio-physical 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, 25.4, 25.11] 25.1. Introduction and Major Conclusions from Previous Assessments Australasia is defined here as lands, territories, offshore waters and oceanic 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 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. Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 • • • • • 25.2. 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 south-west, Kakadu wetlands, coral reefs and sub-Antarctic islands; - water security problems in the Murray-Darling basin, south-western Australia and eastern New Zealand; - potentially large risks to coastal development in south-eastern Queensland and in New Zealand from Northland to the Bay of Plenty. 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 belt 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 Nino-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 Interdecadal 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 (AR5 WGI Ch14) 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 CMIP3 models and SRES scenarios, but CMIP5 model results are considered where available (see also AR5 WGI Chap 14 & Atlas; WGII 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-modelled past and projected future annual average surface temperatures are shown in Figures 25-1 and 25-2. For further details see WGI 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 north-western Australia since the 1950s, the autumn/winter decline since 1970 in south-western Australia and, since the 1990s, in south-eastern 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 north-east 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 Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 confidence in south-western Australia. For New Zealand, annual average rainfall is projected to decrease in the north-eastern 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 Atlas, AI.7071). Examples of the magnitude of projected annual change from 1990 to 2090 (percent model mean change +/intermodel standard deviation) under RCP8.5 from CMIP5 are -20±13% in south-western 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, e.g. 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 south-east 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). Other 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 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 (see Table 25-1). [INSERT FIGURE 25-1 HERE Figure 25-1: Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression [WGI AR5 Figures SPM.1 and 2.21]. (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression [WGI AR5 Figure SPM.2]. For observed temperature and precipitation, trends have been calculated where sufficient data permits 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 change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, 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, less than the baseline variability in >66% of models. (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 21-3 and CC-RC].] [INSERT FIGURE 25-2 HERE Figure 25-2: Observed and simulated variations in past and projected future annual average near-surface air temperature over land areas of Australia (left) and New Zealand (right). Black lines show various estimates from observational measurements. Shading denotes the 5-95 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 ‘RCP2.6’ emissions scenario (63), and the ‘RCP8.5’ (63). Data are anomalies from the 1986-2005 average of the individual observational data (for the observational time series) or of the corresponding historical allforcing simulations. Further details are given in Box 21-3.] Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT TABLE 25-1 HERE 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 Chapters 3, 6 (ocean changes, including acidification), 11, 12 (projections), 13 (sea level) and 14 (regional climate phenomena).] 25.3. Socio-Economic 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 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, GDP grew by an average of 3.2% p.a. in Australia and 2.4% p.a. 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 2.5-3.5% p.a. in Australia and about 1.9% p.a. 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 31-43 million by 2056 and 34-62 million by 2101 (ABS, 2008, 2013); New Zealand’s population from 4.4 million in 2011 to 5.1-7.1 million by 2061 (SNZ, 2012a). The number of people aged 65 and over is projected to almost double in the next two 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, 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; 25.3.2). Indigenous peoples constitute about 2.5% and 15% of the Australian and New Zealand populations, respectively, but in Australia, 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 socio-economic status and social inclusion could strongly influence their future adaptive capacity (see 25.8.2). 25.3.2. Use and Relevance of Socio-Economic Scenarios in Adaptive Capacity/Vulnerability Assessments Demographic, economic and socio-cultural 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, e.g. 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). Socio-economic 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 Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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, 15, 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 Australia, 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 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 (DIICCSRT, 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 organisations 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 et al., 2012a; Johnston et al., 2013; Kuruppu et al., 2013; see also Chapters 10, 16). This creates challenges for achieving holistic societal outcomes (see also 25.7-25.9). Several recent policy initiatives in Australia, while responding to broader socio-economic 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 1997-2009, the Victorian bushfires 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). Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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, 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., 2014). 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, 2012; KCDC, 2012; Lawrence et al., 2013b). Table 25-2 summarises 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 (high agreement, robust evidence; Britton, 2010; Barnett et al., 2013; Lawrence et al., 2013b; Mukheibir et al., 2013; Webb et al., 2013; see also Chapter 16). [INSERT TABLE 25-2 HERE Table 25-2: Constraints and enabling factors for institutional adaptation processes in Australasia.] 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 resourcelimited 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 the uncertain and dynamic nature of climate risk (McDonald, 2010; Stafford-Smith et al., 2011; Gorddard et al., 2012; McDonald, 2013; Reisinger et al., 2014). 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 prioritise 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 emphasise 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). Realising widespread and equitable participation is challenging where policies are complex, debates polarised, 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 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). Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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. _____ START BOX 25-1 HERE _____ 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 centres and infrastructure (see 25.3). Under a high emissions scenario (RCP8.5), 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 metre during the 21st century (AR5 WGI 13.5). Local case studies in New Zealand (Fitzharris, 2010; Reisinger et al., 2014) 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 over A$226 billion of assets, including up to 274,000 residential and 8,600 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 (AR5 WGI 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., 2014). 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 on-going protection and development, which limit options for more transformational responses such as accommodation and retreat (high agreement, medium evidence; Gorddard et al., 2012; Barnett et al., 2013; Fletcher et al., 2013; McDonald, 2013). Strategic regional-scale planning initiatives in rapidly growing regions, like south-east 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., 2014). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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). In addition to raising minimum floor levels and creating coastal set-backs 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., 2014). _____ END BOX 25-1 HERE _____ 25.4.3. Psychological and Socio-cultural 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 behaviourally (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 behaviour (high agreement, medium evidence; 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). 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 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- Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 behavioural 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 behavioural 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 bio-physical and economic characterisations 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 south-eastern Australia resulted in low irrigation water allocations, severe water restrictions in urban centres, 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 resources in far south-eastern and far south-west Australia will decline (high confidence; by 0-40% and 20-70%, 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 2–3 times larger percent change in mean annual streamflow (Chiew, 2006; Jones et al., 2006). This can vary, however, with unprecedented declines in flow in far south-eastern Australia in the 1997–2009 drought (Cai and Cowan, 2008; Potter and Chiew, 2011; Chiew et al., 2013). Higher temperatures and associated evaporation, tree re-growth following more frequent bushfires (Kuczera, 1987; Cornish and Vertessy, 2001; Marcar et al., 2006; Lucas et 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 CO2 environment (Betts et al., 2007; Donohue et al., 2009; McVicar et al., 2010). [INSERT FIGURE 25-4 HERE 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 (millimetres; 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).] Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 north-east 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, Waimakariri, Rangitata) are projected to increase by 5-10 % (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; McCallum et al., 2010; Crosbie et al., 2012). 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 degrade water quality, particularly through increased material washoff following bushfires and floods (Box 25-6, Box 25-8). 25.5.3. Adaptation The 1997–2009 drought in south-eastern 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. _____ START BOX 25-2 HERE _____ Box 25-2. Adaptation through Water Resources Policy and Management in Australia Widespread drought and projections of a drier future in south-eastern and far south-west 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 recognises 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 on-going revisions to allocation plans and time lags to observable impacts. Rural water reform in south-eastern Australia, focused on the Murray-Darling Basin, is currently being implemented. The Murray-Darling Basin Plan (MDBA, 2011, 2012) will return 2750 GL/year 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 A$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 optimise 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 Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 south-eastern 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 realised (25.5.2, Figure 25-4), agriculture and ecosystems across south-eastern and south-western Australia would be threatened even with comprehensive adaptation (see 25.6.1, 25.7.1, 25.7.2; Connor et al., 2009; Kirby et al., 2013). Climate change and population growth are the two major factors that influence water planning in Australian capital cities. In Melbourne, for example, planning has centred on securing new supplies that are more resilient to major climate shocks; increasing use of alternative sources like 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 desalinisation plant with a 150 GL/year capacity (about one third of the current demand), following the lead of Perth where a desalinisation 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 desalinisation 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). Desalinisation plants can be maladaptive due to 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). _____ END BOX 25-2 HERE _____ 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 CO2, fire management, grazing and land-use. The 1997-2009 drought had Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 overallocation, 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 south-east Australia (Mac Nally et al., 2009), savannah trees in north-east Australia (Fensham et al., 2009; Allen et al., 2010), mediterranean-type eucalypt forest in southwest Western Australia (Matusik et al., 2013), and, eucalypts in sub-alpine regions in Tasmania (Calder and Kirkpatrick, 2008). Mass dieoffs of flying foxes and cockatoos have been observed during heatwaves (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 modelling (SDM) consistently indicates future range contractions for Australia’s native species even assuming optimistic rates of dispersal, e.g. Western Australian Banksia spp. (Fitzpatrick et al., 2008), koalas (Adams-Hosking et al., 2011), northern macropods (Ritchie and Bolitho, 2008), native rats (Green et al., 2008b), 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 due to 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 savannas subject to changed fire regimes (Laurance et al., 2011); inland freshwater and groundwater systems subject to drought, overallocation 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; Yates et al., 2010b) and tropical and sub-tropical rainforests 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 on-going 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 increase 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 Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 colonisation 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). Climate 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 (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, and 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 ~8.1 million km2, excluding the Australian Antarctic Territory (Richardson and Poloczanska, 2009). New Zealand has ~18,000 km of coastline, spanning subtropical to subAntarctic waters, and the world's fifth largest Exclusive Economic Zone at 4.2 million km2 (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 25.3) and Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 <1% of the land mass but may account for 39% of Australia’s average national annual carbon burial (estimated total: 466 millions tonnes CO2-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 ~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 3.8, 6.44). Recently observed changes in marine systems around Australia are consistent with warming oceans (high confidence; Box 25-3). Examples include 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 south-east 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 over $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 sealevel 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 Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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; 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, e.g. loss of 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 (GBR) as highly vulnerable to both warming and acidification (Hennessy et al., 2007). Recent observations of bleaching (GBRMPA, 2009a) and reduced calcification in both the GBR 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 CO2 concentrations in excess of 450-500 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, 2011). 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 sub-tropical 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 and 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 colonisation 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). Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 colonisation 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 colonisation to maintain production in the face of declining recruitment may also be possible for some high value species, and has been trialled for the southern rock lobster (Green et al., 2010a). 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). _____ START BOX 25-3 HERE _____ Box 25-3. Impacts of a Changing Climate in Natural and Managed Ecosystems Observed changes in species, and in natural and managed ecosystems (25.6.1, 25.6.2, 25.7.2) provide multiple lines of evidence of the impacts of a changing climate1. Examples of observations published since the AR4 are shown in Table 25-3. [INSERT TABLE 25-3 HERE Table 25-3: Examples of detected changes in species, natural and managed ecosystems, consistent with a climate change1 signal, published since the AR4. Confidence 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. Confidence in the role of climate as 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 insufficient to explain the observed change.] [FOOTNOTE 1: Consistent with the IPCC definition, a change in climate refers to any statistically detectable signal; it does not necessarily imply a human cause. See Glossary, Table 25-1 and 25.2.] _____ END BOX 25-3 HERE _____ 25.7. Major Industries 25.7.1. Production Forestry Australia has about 149 Mha forests, including woodlands. Two Mha are plantations and 9.4 Mha multiple-use native forests, and forestry contributes around $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). 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. Modelled projections are based on ecophysiological responses of forests to CO2, water and temperatures. In Australia, potential changes in water availability will be most important (very high confidence; e.g. Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 reviews by Battaglia et al., 2009; Medlyn et al., 2011b). Modelling future distributions or growth rates indicate that plantations in south-west 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 fertilisation from increasing CO2 (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 CO2 fertilisation may offset this (medium confidence; Kirschbaum et al., 2012). Modelling studies are limited by their reliance on key assumptions which are difficult to verify experimentally, e.g. the degree to which photosynthesis remains stimulated under elevated CO2 (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 CO2, 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 towards 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 CO2 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 and 18; Webb et al., 2012a; Darbyshire et al., 2013). Because the region is a major exporter – providing, for example, over 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 1980-99 baseline) is expected to be a 4% reduction in gross value of the beef, sheep and wool sector (McKeon et al., 2008). Dairy output is projected to decline in all regions of Australia other than Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Tasmania under a 1°C increase by 2030 (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 CO2 fertilisation and nitrogen feedbacks. Studies modelling seasonal changes in fodder supply show greater sensitivity in animal production to climate change and elevated CO2 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) modelled 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 1960-2008 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 inter-annual 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 et al., 2012b). The ‘water-sparing’ effect of elevated CO2 (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 impacts of elevated CO2 on forage production, quality, nutrient cycling and water availability remains the major uncertainty in modelling 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 CO2 have added to this uncertainty. New Zealand agroecosystems are subject to erosion processes strongly driven by climate; greater certainty in projections of rainfall, particularly storm frequency, are needed to better understand 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 CO2 at two sites with different temperatures have shown a wide range in the response of current wheat cultivars (Fitzgerald et al., 2010). Modelling 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 are projected to result in increased wheat yields in high rainfall areas such as southern Victoria under climate change and maintenance of current yields in some areas expected to be drier (e.g. north-western 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 modelling 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 CO2 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, Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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). Observed trends and modelling 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 et al., 2012a). 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 CO2 impacts on weeds, pests and diseases are highly uncertain (see Box 25.4). Future performance of currently effective plant resistance mechanisms under elevated CO2 and temperature is particularly important (Melloy et al., 2010; Chakraborty et al., 2011) as is the future efficacy of widely used biocontrol, i.e. 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 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 (high agreement, robust evidence; Garnaut, 2008; Quiggin et al., 2010; Qureshi et al., 2013b), e.g. in one study by 12-44% to 2030 and 49-72% 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 over 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 (600-1000 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). _____ START BOX 25-4 HERE _____ 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. [INSERT TABLE 25-4 HERE 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).] Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ END BOX 25-4 HERE _____ _____ START BOX 25-5 HERE _____ 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 25.4.2, 25.6.1, 25.7.2, 25.7.3, Box 25-2, Box 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 (see 25.7, 25.8; Kouvelis et al., 2010; Balston et al., 2012). 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 (high agreement, robust evidence; Alston, 2007; Edwards and Gray, 2009; Alston, 2010, 2012; Hanigan et al., 2012; see also Box CC-GC). Positive social change also occurred, however, including increased social capital through interaction with community organisations (Edwards and Gray, 2009). While social and cultural changes have the potential to undermine the adaptive capacity of communities (Smith 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 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 2012-13 drought in New Zealand is estimated to have reduced national GDP by 0.3-0.6% 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), e.g. 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; 25.7.2), with positive or negative implications for the wider communities in origin and destination regions (Kiem and Austin, 2012). Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Although stakeholders within rural communities differ in their vulnerabilities and adaptive capacities, they are bound by similar dependence upon 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). _____ END BOX 25-5 HERE _____ _____ START BOX 25-6 HERE _____ 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, burnt over 3,500 km2, caused 173 deaths, destroyed over 2,000 buildings and caused damages of A$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 savannas 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 CO2 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 south-east and south-west (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 et al., 2011a). 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 centre 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; Pearce et al., 2011), although changes in management to date show little evidence of being driven by climate change. _____ END BOX 25-6 HERE _____ Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 25-54 million tonnes and led to A$5-9bn 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; Hodgkinson et al., 2010b). 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. On-going challenges include competition for energy and water, climate change scepticism, 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). 25.7.4. Energy Supply, Demand, and Transmission Energy demand is projected to grow by 0.5-1.3% 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 ~33-50% 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 5-10% 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 power generation (MED, 2011) would benefit from projected increases in mean westerly winds but face increased risk of damages and shut-down during extreme winds (Renwick et al., 2009). Climate warming would reduce annual average peak electricity demands by 1-2% per °C 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 black-outs (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 A$2.5 billion to 2015, with more than half to meet increasing demand for air conditioning and the remainder to increase resilience to climaterelated hazards; underground cabling would reduce bushfire risk but has large investment costs that are not included above (Parsons Brinkerhoff, 2009). Decentralised ownership of assets constitutes a significant adaptation constraint Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (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 2.6-4% 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 A$51.4 billion (Oxford Economics, 2009). Most Australasian tourism is exposed to climate variability and change (see 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 A$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 A$70 million loss in 2008 due to reduced visitor days (TRA, 2010). 25.7.5.1. Projected Impacts Future impacts on tourism have been modelled for several Australian destinations. The Great Barrier Reef is expected to degrade under all climate change scenarios (25.6.2, 30.5, Box CC-CR), reducing its 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 modelling 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 analogue-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, due to uncertain future tourist behaviour (Scott et al., 2012; also 25.9.2). 25.7.5.2. Adaptation Both New Zealand and Australia have formalised 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 A$100 million in capital investment into 700 snow guns and 2.5-3.3 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 due to limited Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 understanding of future changes in tourism and community preferences (Scott et al., 2012), including the flow-on effects of changing travel behaviour and tourism preferences in other world regions (see 25.9.2). 25.8.1. Human Health 25.8.1.1. Observed Impacts Life expectancy in Australasia is high, but shows substantial ethnic and socio-economic inequalities (Anderson et al., 2006). Mortality increases in hot weather in Australia (high agreement, robust evidence; Bi and Parton, 2008; Vaneckova et al., 2008) with air pollution exacerbating this association. The last four 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 heatwave 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; Tong et al., 2010b). For example, during the heatwave in January and February 2009 in south-eastern Australia (BoM, 2009), total emergency 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 heatwaves (1993-2006; Hansen et al., 2008). Mortality attributed to mental and behavioural 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 25.4.3). 25.8.1.2. Projected Impacts Projected increases in heatwaves (Figure 25-5) will increase heat-related deaths and hospitalizations, especially among the elderly, compounded by population growth and ageing (high confidence; Bambrick et al., 2008; 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 details 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 hospitalisations 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, 205,000-335,000 new cases of bacterial gastroenteritis by 2050, and 239,000-870,000 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 of the combined effects of climate change and socio-economic 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 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 socio-economic factors, especially domestic water-storage may have a more Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 socio-economic factors, and thus will vary substantially among different coastal regions (Naish et al., 2013). The effects of climate change combined with frequent travel 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; 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). [INSERT FIGURE 25-5 HERE 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 ~1990 (based on available meteorological station data for the period 1975-2004), ~2050 and ~2100. Bar charts show the change in population heat exposure, expressed as persondays 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 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).] 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 healthcare 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 behaviours (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 Box 25-6, Box 25-8). A review of the southern Australian heatwave 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 heatwave 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). Socio-economic 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, assessments have not specifically addressed risks to Indigenous people (e.g. Guillaume et al., 2010). In other Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 regions, all remote, there is limited empirical evidence of vulnerability (Maru et al., 2012). However, there is high agreement and medium evidence 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; NNTT, 2013). There is high agreement but limited evidence that natural resource dependence (e.g. Bird et al., 2005; Gray et al., 2005a; 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 et al., 2010b; TSRA 2010). Little adaptation of Indigenous communities to climate change is apparent to date (but see 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 et al., 2013b; 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 et al., 2010c; Petheram et al., 2010) and developing inter-cultural communication tools (Woodward et al., 2012; Leonard et al., 2013b). 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). The 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 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., Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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). Māori knowledge of environmental processes and hazards (King et al., 2005; King et al., 2007) as well as strong social-cultural networks are vital for adaptation and on-going 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). _____ START BOX 25-7 HERE _____ 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 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 A$3 billion p.a. (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). Normalised insured losses in Australia show no significant trend from 1967 to 2006 (Crompton and McAneney, 2008; Crompton et al., 2010; Table 10.4), though this conclusion rests on a simplified accounting of population growth and may also reflect improved building codes and early warning systems (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, e.g., through unaffordability or unavailability of cover in areas at highest risk (IAA, 2011b, a; 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), e.g. 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-subsidised 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., 2014; see also 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). _____ END BOX 25-7 HERE _____ _____ START BOX 25-8 HERE _____ 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 Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 infrastructure of AUD$5-6 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 modelling in New Zealand (Gray et al., 2005b; 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 5–10% 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 (Box 25.9; Howe et al., 2005; Skinner, 2010; WCC, 2010). 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; 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 realised (high confidence; Lawrence and Allan, 2009; DERM et al., 2010; Glavovic et al., 2010; Wilby and Keenan, 2012; Lawrence et al., 2013a). _____ END BOX 25-8 HERE _____ 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). Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ START BOX 25-9 HERE _____ 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 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 and 10, 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 summarises some adaptation options, co-benefits and constraints on their adoption in Australasia. 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 (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; Britton, 2010; Norman, 2010; Abel et al., 2011; Lawrence et al., 2013a; McDonald, 2013; Palutikof et al., 2013; Reisinger et al., 2014). [INSERT TABLE 25-5 HERE Table 25-5: Examples of co-beneficial 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.] _____ END BOX 25-9 HERE _____ 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 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). Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT TABLE 25-6 HERE 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 cause negative impacts or have co-benefits with impacts or responses in another sector, or with another type of response in the same sector.] 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 (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 (25.7.4, Box 25-9). Table 25-7 gives further 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. [INSERT TABLE 25-7 HERE Table 25-7: Examples of interactions between adaptation and mitigation measures (green rows denote synergies where multiple benefits may be realized, orange rows denote potential tradeoffs and conflicts; grey row gives an example of complex, mixed interactions). The primary goal may be adaptation or mitigation.] _____ START BOX 25-10 HERE _____ 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 towards 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 modelled 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). Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ END BOX 25-10 HERE _____ 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. Modelling 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 due to climate change, especially towards the end of the 21st century (Gunasekera et al., 2008). These conclusions, however, merit only medium confidence, because they rely on simplified assumptions 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 0.62.3% 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 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). Climate change has the potential to change migration flows within Australasia, particularly due to coastal changes (e.g. from the Torres Straits islands to mainland Australia), although reliable estimates of such movements do not yet exist (see 12.4; Green et al., 2010b; 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 et al., 2008a; 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 (see 12.4, 12.5; Dupont, 2008; Pearman, 2009). 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 (high agreement, robust evidence; 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 Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 mitigation and non-mitigation scenarios (see also 19.7). Literature quantifying these benefits for Australasia has increased since the AR4 but remains very sparse. Economy-wide net costs for Australia are modelled 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 CO2-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 implementation 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, e.g. for irrigated agriculture in the Murray-Darling Basin (Quiggin et al., 2008; Quiggin et al., 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 25.7, 25.8) 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, e.g. heat and cold 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 CO2 fertilization (Kirschbaum et al., 2011a; 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 riskmanagement perspective. [INSERT TABLE 25-8 HERE 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 longterm (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.] One 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 Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 change (high confidence). A second set of key risks (increase in flood risk, water scarcity, heat waves and wild fire) 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 25.5.2) or rapid sea level rise approaching 1 m or more by 2100 (see Box 25-2 and WGI 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 (CSIRO, 2011; Leonard et al., 2013a), 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 (25.7.1, 25.7.2, 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 set of assumptions, models, and data and which, 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 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 (25.6.1, 25.6.2). Maintenance of key ecosystem functions and services requires a radical reassessment of conservation values and practices related to assisted colonisation 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 Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 (high agreement, medium evidence; 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, -2, 5-9; 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 due to uncertainty about the future will not necessarily minimize costs or ensure adequate flexibility for future responses, 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 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 CO2, 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 CO2 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 socio-economic characteristics. Research into psychological, economic, social and cultural dimensions of vulnerability, adaptive capacity and underpinning values remains limited and poorly integrated with bio-physical 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. While 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. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate change impacts, adaptation and mitigation responses in other world regions will affect Australasia, but our understanding of this remains 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 socio-economic changes into studies of local vulnerability and adaptation options. Frequently Asked Questions FAQ 25-1: How can we adapt to climate change if projected future changes remain uncertain? [to be inserted at end of Section 25.4.2] Many existing climate change impact assessments in Australia and New Zealand focus on the distant future (2050 to 2100). 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 recognises 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, and 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. 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 which would cope only with a limited sub-set of possible climate futures. This is sometimes referred to as maladaptation. Changing pathways – for example, shifting from on-going 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. [INSERT FIGURE 25-3 HERE 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.] FAQ 25-2: What are the key risks from climate change to Australia and New Zealand? [to be inserted at end of Section 25.10.3] Our assessment identifies eight key regional risks from climate change. Some impacts, especially on ecosystems, are 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 even 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 Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 most parts of Australia and New Zealand; constraints on water resources from reducing rainfall in southern Australia; increased health 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 south-eastern and far south-western 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 materialise 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. References ABARES, 2011: Australian mineral statistics 2011, March Quarter 2011. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra, 42 pp. ABARES, 2012: Australia’s Forests at a Glance 2012. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra, 83+vii pp. Abbs, D.J., 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. CSIRO, Aspendale, 83 pp. Abbs, D.J., 2012: The impact of climate change on the climatology of tropical cyclones in the Australian region. Climate Adaptation Flagship Working Paper No 11. CSIRO, Aspendale, 24 pp. Abdo, D.A., D.A. Bellchambers, L.M. Evans, 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. Abel, N., R. Gorddard, B. Harman, A. Leitch, J. Langridge, A. Ryan, 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, Canberra, 100 pp. ABS, 2009: Experimental estimates and projections, Aboriginal and Torres Strait Islander Australians, Catalogue 3238.0. Australian Bureau of Statistics, Canberra, 76 pp. ABS, 2010a: Tourism satellite account 2009-10. Catalogue 5249.0. Australian Bureau of Statistics, Canberra, 50 pp. ABS, 2010b: Population characteristics, Aboriginal and Tores Trait Islander Australians. Catalogue No. 4713.0. Australian Bureau of Statistics, Canberra, 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, Canberra, 107 pp. ABS, 2012a: Water Account Australia 2009-10. Catalogue 4610.0. Australian Bureau of Statistics, Canberra, 93 pp. ABS, 2012b: Gross Value of Irrigated Agricultural Production, 2010-11. Catalogue No. 4610.0.55.008 (online at www.abs.gov.au). Australian Bureau of Statistics, Canberra. ABS, 2012c: Yearbook Australia 2012. Catalogue No. 1301.0. Australian Bureau of Statistics, Canberra, 885 pp. ABS, 2012d: Australian National Accounts: National Income, Expenditure and Product, Jun 2012. Table 5206.0. Australian Bureau of Statistics, Canberra, 81 pp. ABS, 2013: Australian Demographic Statistics. Australian Bureau of Statistics, Canberra, 108 pp. Ackerley, D., R. Bell, B. Mullan, H. McMillan, 2013: Estimation of regional departures from global-average sealevel 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, ACT Government, Canberra, 94 pp. Adams, M., P. Attiwill, 2011: Burning Issues: Sustainability and Management of Australia's Southern Forests. CSIRO Publishing, Collingwood, Vic, 144 pp. Adams-Hosking, C., H.S. Grantham, J.R. Rhodes, C. McAlpine, P.T. Moss, 2011: Modelling climate-changeinduced shifts in the distribution of the koala. Wildlife Research, 38(2), 122-130. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Agho, K., G. Stevens, M. Taylor, M. Barr, 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, Canberra, 73 pp. Akompad, D.A., P. Bi, S. Williams, J. Grant, I.A. Walker, M. Augoustinous, 2013: Heat waves and climate change: applying the health belief model to identify predicotrs of risk preception and adapative behaviours in Adelaide, Australia. International Journal of Environmental Reserach and Public Health, 10, 2164-2184. Alexander, L.V., P. Hope, D. Collins, B. Trewin, A. Lynch, N. Nicholls, 2007: Trends in Australia's climate means and extremes: a global context. Australian Meteorological Magazine, 56(1), 1-18. Alexander, L.V., 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., X.L. Wang, H. Wan, B. Trewin, 2011: Significant decline in storminess over southeast Australia since the late 19th century. Australian Meteorological and Oceanographic Journal, 61, 23-30. Alexander, K.S., A. Ryan, 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. 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, 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., D.J. Karoly, 2013: A climatology of Australian severe thunderstorm environments 1979–2011: interannual variability and ENSO influence. International Journal of Climatology, advance online, 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, 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, ANU, Canberra, 65 pp. Anderson, I., S. Crengle, M. Leialoha Kamaka, T.-H. Chen, N. Palafox, L. Jackson-Pulver, 2006: Indigenous health in Australia, New Zealand, and the Pacific. The Lancet, 367(9524), 1775-1785. Anderson, B., W. Lawson, I. Owens, 2008: Response of Franz Josef Glacier Ka Roimata o Hine Hukatere to climate change. Global and Planetary Change, 63(1), 23-30. Anthony, K.R.N., D.I. Kline, G. Diaz-Pulido, S. Dove, O. Hoegh-Guldberg, 2008: Ocean acidification causes bleaching and productivity loss in coral reef builders. Proceedings of the National Academy of Sciences, 105(45), 17442-17446. ANU, 2009: Implications of climate change for Australia’s World Heritage properties: A preliminary assessment. Report by the Fenner School of Environment and Society, Australian National University. Department of Climate Change, Department of the Environment, Water, Heritage and the Arts, Canberra, 207 pp. Arkema, K.K., G. Guannel, G. Verutes, S.A. Wood, A. Guerry, M. Ruckelshaus, P. Kareiva, M. Lacayo, J.M. Silver, 2013: Coastal habitats shield people and property from sea-level rise and storms. Nature Climate Change, advance online, DOI: 10.1038/nclimate1944. Ashworth, P., T. Jeanneret, J. Gardner, H. Shaw, 2011: Communication and climate change: What the Australian public thinks. Report EP112769. CSIRO, Canberra, 68 pp. 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, 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, 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, 71 pp. Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ATSISJC, 2009: Native Title Report 2008. Aboriginal and Torres Strait Islander Social Justice Commissioner, Australian Human Rights Commission, Sydney, 422 pp. Auckland Council, 2012: The Auckland Plan. Auckland Council, Auckland, 373 pp. August, S.M., 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, 365 pp. Australian Treasury, 2010: Australia to 2050: future challenges. The 2010 Intergenerational Report. Australian Treasury, Canberra, 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, University of Technology, Sydney, 70 pp. Baisden, W.T., E.D. Keller, L. Timar, D. Smeaton, A. Clark, A. Ausseil, W. Power, W. Zhang, 2010: New Zealand’s Pasture Production in 2020 and 2050. Consultancy Report 2010/154 by GNS Sciences for the Ministry of Agriculture and Forestry. Ministry for Primary Industries, Wellington, New Zealand, 68 pp. Ballinger, J., B. Jackson, A. Reisinger, 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, NZ, 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 of South Australia, Adelaide, 61 pp. Balston, J.M., J. Kellett, G. Wells, S. Li, A. Gray, I. Iankov, 2012: Quantifying the costs of climate change on local government assets. National Climate Change Adaptation Research Facility, Gold Coast, Qld, 219 pp. Bambrick, H.J., K. Dear, R. Woodruff, I. Hanigan, A.M. Michael, 2008: The impacts of climate change on three health outcomes. Paper prepared for Garnaut Climate Change Review. Garnaut Climate Change Review, Canberra, 47 pp. Banfai, D.S., 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, 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., Nel, P., Righarts, M. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, NZ, pp. 59-70. Barnett, J., S. O'Neill, 2010: Maladaptation. Global Environmental Change, 20(2), 211-213. Barnett, J., E. Waters, S. Pendergast, A. Puleston, 2013: Barriers to adaptation to sea-level rise. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 85 pp. Barton, A.B., 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, 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, 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, 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, W00H02. Bates, B.C., S. Bunn, P. Baker, M. Cox, A. Hopkins, B. Humphreys, S. Lakes, G. Willgoose, B. Young, 2011: National Climate Change Adaptation Research Plan for Freshwater Biodiversity. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 68 pp. Battaglene, S.C., C. Carter, A.J. Hobday, V. Lyne, 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, University of Tasmania, Hobart, 84 pp. Battaglia, M., J. Bruce, C. Brack, T. Baker, 2009: Climate Change and Australia's Plantation Estate: Analysis of Vulnerability and Preliminary Investigation of Adaptation Options. Forest & Wood Products Australia Limited, Melbourne, Vic, 130 pp. Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Baum, S., S. Horton, D.L. Choy, 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, 76 pp. Baynes, T., A. Herr, A. Langston, H. Schandl, 2012: Coastal Climate Risk Project Milestone 1 Final report. Department of Climate Change and Energy Efficiency, Canberra, 104 pp. Becken, S., 2011: A Critical Review of Tourism and Oil. Annals of Tourism Research, 38(2), 359-379. Becken, S., R. Clapcott, 2011: National tourism policy for climate change. Journal of Policy Research in Tourism, Leisure and Events, 3(1), 1-17. Bedford, R., 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. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, pp. 89-134. Bedford, R., J. Campbell, 2013: Migration and climate change in Oceania. In: People on the move in a changing climate [Piguet, E., Laczko, F. (eds.)]. Springer, Dordrecht (ISBN 978-94-007-6984-7). Beebe, N.W., R.D. Cooper, P. Mottram, A.W. Sweenet, 2009: Australia's dengue risk driven by human adaptation to climate change. PLOS Neglected Tropical Diseases, 3(5), e429. Beentjes, M.P., 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., 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, 46S-53S. Béguin, A., S. Hales, J. Rocklöv, C. Åström, V.R. Louis, 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, P. Vowles, 2010: Challenges and Choices: Modelling New Zealand’s Long-term Fiscal Position. Working Paper 10/01. New Zealand Treasury, Wellington, NZ, 85 pp. Bennett, C.M., K.B.G. Dear, A.J. McMichael, 2013: Shifts in the seasonal dsitribution of deaths in Australia, 19682007. International Journal of Biometeorology, advance online, DOI: 10.1007/s00484-013-0663-x. Bergin, A., J. Townsend, 2007: A change in climate for the Australian defence force. Special Report 2007-7. Australian Strategic Policy Institute, Canberra, 11 pp. Berry, S., J. Vella, 2010: Planning Controls and Property Rights – Striking the Balance. Roadshow 2010. Resource Management Law Association, Auckland, 66 pp. Berry, H.L., J.R.A. Butler, C.P. Burgess, U.G. King, K. Tsey, Y.L. Cadet-James, C.W. Rigby, B. Raphael, 2010: Mind, body, spirit: co-benefits for mental health from climate change adaptation and caring for country in remote Aboriginal Australian communities. NSW Public Health Bulletin, 21(5-6), 139-45. 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, 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 by Sean Bevin, Economic Solutions LTD, Napier. Ministry of Agriculture and Forestry, Wellington, 20 pp. Bi, P., 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., 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, 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, 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. Department of Climate Change and Energy Efficiency, Canberra, 229 pp. Bohensky, E., J.R.A. Butler, R. Costanza, I. Bohnet, A. Delisle, K. Fabricius, M. Gooch, I. Kubiszewski, G. Lukacs, P. Pert, 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, Canberra, 11 pp. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 BoM, 2013: Special Climate Statement 43 – extreme heat in January 2013. Bureau of Meteorology, Canberra, 23 pp. Bond, N., P. Lake, A. Arthington, 2008: The impacts of drought on freshwater ecosystems: an Australian perspective. Hydrobiologia, 600(1), 3-16. Bond, N., J. Thomson, P. Reich, J. Stein, 2011: Using species distribution models to infer potential climate changeinduced range shifts of freshwater fish in south-eastern Australia. Marine and Freshwater Research, 62(9), 1043-1061. Booth, T.H., M.U.F. Kirschbaum, M. Battaglia, 2010: Forestry. In: Adapting Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 137-152. Booth, K., 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. Botterill, L.C., 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., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 100-114. Bourdôt, G.W., S.L. Lamoureaux, M.S. Watt, L.K. Manning, D.J. Kriticos, 2012: The potential global distribution of the invasive weed Nassella neesiana under current and future climates. Biological Invasions, 14(8), 15451556. Bowman, D., B.P. Murphy, 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, N.S. Sodhi, 2010: Evaluating the Relative Environmental Impact of Countries. Plos One, 5(5), e10440. 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, 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. Australian Greenhouse Office and Department of the Environment and Water Resources, Canberra, 81 pp. Bright, J., H. Rutter, J. Dommisse, R. Woods, A. Tait, B. Mullan, J. Hendrix, J. Diettrich, 2008: Projected Effects of Climate Change on Water Supply Reliability in mid-Canterbury. Report C08120/1 by Aqualine Research Ltd. Ministry of Agriculture and Forestry, Wellington, 45 pp. 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), Wellington, 51 pp. Britton, E., S. Hales, K. Venugopal, 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, 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., J. Dahm, H. Rouse, T. Hume, R. Bell, P. Blackett, 2011: Coastal Adaptation to Climate Change: Pathways to Change. NIWA, Auckland, 106 pp. Brockerhoff, E.G., H. Jactel, J.A. Parrotta, C.P. Quine, 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, 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, 2861. Brown, R.R., M. Farrelly, 2009: Challenges ahead: social and institutional factors influencing sustainable urban stormwater management in Australia. Water Science and Technology, 59(4), 653-660. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Brown, R.R., M. Farrelly, 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, L. Thompson, 2011: Hunter & Central Coasts New South Wales – Vulnerability to climate change impacts. Department of Climate Change and Energy Efficiency, Canberra, 149 pp. Bruno, J.F., E.R. Selig, K.S. Casey, C.A. Page, B.L. Willis, C.D. Harvell, H. Sweatman, 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, 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, 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, S.A. Setterfield, 2011: Is Australia ready for assisted colonization? Policy changes required to facilitate translocations under climate change. Pacific Conservation bBiology, 17(3), 259-269. Burgess, C.P., F.H. Johnston, H. Berry, J. McDonnell, D. Yibarbuk, C. Gunabarra, A. Mileran, 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. Burgette, R.J.W., C.S, J.A. Church, N.J. White, P. Tregoning, 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, advance online, DOI: 10.1093/gji/ggt131. Burroughs, 2010: Sustainable housing for Indigenous populations in remote regions: A case study from Nguiu, Northern Australia. In: Keynote Lectures: CESB 10: Central Europe towards Sustainable Building. From Theory to Practice, Czech Technical University, Prague, Czech Republic. pp 8. Burton, P., J. Mustelin, 2013: Planning for climate change: is greater public participation the key to success? Urban Policy and Research, advance online, DOI: 1080/08111146.2013.778196. Butcher, 2009: Regional and national impacts of the 2007–2009 drought. Report by Butcher Partners Ltd. Ministry of Agriculture and Forestry, Wellington, 33 pp. Buultjens, J., D. Gale, 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, Q., Connell, D. (eds.)]. ANU E-Press, Canberra, pp. 385-398. Cai, W., 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. Cai, W., T. Cowan, 2008: Evidence of impacts from rising temperature on inflows to the Murray-Darling Basin. Geophysical Research Letters, 35, L07701. Cai, W., T. Cowan, M. Raupach, 2009a: Positive Indian Ocean Dipole events precondition southeast Australia bushfires. Geophysical Research Letters, 36(19), L19710. Cai, W., A. Sullivan, T. Cowan, 2009b: Climate change contributes to more frequent consecutive positive Indian Ocean Dipole events. Geophysical Research Letters, 36(23), L23704. Cai, W., P. van Rensch, 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, T. Cowan, 2012: The 2011 southeast Queensland floods: a confirmation of a negative Pacific Decadal Oscillation phase? Geophysical Research Letters, 39(8), L08702. Calder, J.A., 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., 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, H. Cleland, 2009: Black Saturday: the immediate impact of the February 2009 bushfires in Victoria, Australia. The Medical Journal of Australia, 191, 11-16. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Campbell, A., 2008: Managing Australian Landscapes in a Changing Climate: A climate change primer for regional Natural Resource Management bodies. Department of Climate Change, Canberra, 47 pp. Campbell, D., M. Stafford Smith, J. Davies, P. Kuipers, J. Wakerman, M.J. McGregor, 2008: Responding to health impacts of climate change in the Australian desert. Rural and Remote Health, 8(3), 1008. Carey-Smith, T., S. Dean, J. Vial, C. Thompson, 2010: Changes in precipitation extremes for New Zealand: climate model predictions. Weather and Climate, 30, 23-48. Carr, M.K.V., 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, S. Kerr, 2002: A Framework For Engagement Of Māori Landowners In “Carbon Farming” Using Indigenous Forest Regeneration. Landcare Research Contract Report: LC0102/116. Landcare Research, Lincoln, 24 pp. Carter, T.R., 2013: Agricultural impacts: Multi-model yield projections. Nature Climate Change, 3(9), 784-786. Cary, G., D. Lindenmayer, S. Dovers (eds.), 2003: Australia Burning - Fire Ecology, Policy and Management Issues. CSIRO Publishing, Collingwood, Vic, 280 pp. CCC, 2010: Christchurch City Council climate smart strategy 2010-2025. Christchurch City Council, Christchurch, 26 pp. Chakraborty, S., J. Luck, G. Hollaway, G. Fitzgerald, N. White, 2011: Rust-proofing wheat for a changing climate. Euphytica, 179, 19-32. Chambers, L.E., 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, 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, A.C. Wolfaardt, 2013a: Phenological changes in the Southern Hemisphere. Plos One, advance online, DOI: 10.1371/journal.pone.0075514. Chambers, L.E., L. Beaumont, I. Hudson, 2013b: Continental scale analysis of bird migration timing: influences of climate and life history traits; a generalised mixture model clustering approach. Int. J. Biometeorology, advance online, 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, D. Pauly, 2012: Climate-change induced tropicalisation of marine communities in Western Australia. Marine and Freshwater Research, 63(5), 415-427. Chiew, F., 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., 2006: Estimation of rainfall elasticity of streamflow in Australia. Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 51(4), 613-625. Chiew, F., J. Teng, J. Vaze, D.A. Post, J.M. Perraud, D.G.C. Kirono, 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. Chiew, F., I.P. Prosser, 2011: Water and Climate. In: Water: Science and Solutions for Australia [Prosser, I. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 29-46. Chiew, F., W.J. Young, W. Cai, 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, D.A. David, 2013: Observed hydrologic nonstationarity in far south-eastern Australia: implications and future modelling predictions. Stochastic Environmental Research and Risk Assessment, advance online, DOI: 10.1007/s00477-013-0755-5. 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, 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. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chinn, T.J., B.B. Fitzharris, A. Willsman, 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., 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, 30 pp. City of Melbourne, 2009: City of Melbourne Climate Change Adaptation Strategy. Prepared by Maunsell Australia. City of Melbourne, Melbourne, 128 pp. City of Port Phillip, 2010: Greening Port Phillip, an urban forest approach. Port Phillip City Council, St Kilda, Vic, 54 pp. Clark, A., A. Tait, 2008: Drought, Agricultural Production & Climate Change – A Way Forward to a Better Understanding. NIWA Client Report WLG2008-33. National Institute of Water & Atmospheric Research Ltd, Wellington, 71 pp. Clark, A., B. Mullan, A. Porteous, 2011: Scenarios of regional drought under climate change. NIWA Client Report WLG2012-32. National Institute of Water and Atmospheric Research (NIWA), Wellington, 135 pp. Clark, A., R. Nottage, D. Hansford (eds.), 2012: Impacts of Climate Change on Land-based Sectors and Adaptation Options. Stakeholder Report. Ministry of Primary Industries, Wellington, 74 pp. Clarke, H.G., P.L. Smith, 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, 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, P. Harris, 2011: Natural resource management policy and planning in Australia. Synthesis, Literature Review and Workshop Report. HC Coombs Policy Forum-Fenner School of Environment and Society NRM initiative, The Australian National University, Canberra, 37 pp. Clothier, B., A. Hall, 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., Nottage, R. (eds.)]. Ministry for Primary Industries, Wellington, 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, Department of the Prime Minister and Cabinet, Canberra, 27 pp. Cobon, D.H., G.S. Stone, J.O. Carter, J.C. Scanlan, N.R. Toombs, X. Zhang, J. Willcocks, 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, 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., J. Dibden, 2009: Systems in peril: Climate change, agriculture and biodiversity in Australia. IOP Conference Series: Earth and Environmental Science, 8(1), 012013. Collins, D., 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. Connell, D., Q. Grafton (eds.), 2011: Basin Futures: water reform in the Murray-Darling Basin. ANU E-Press, Canberra, 500 pp. Connor, J., J. Schwabe, D. King, K. D., M. Kirby, 2009: Impacts of climate change on lower Murray irrigation. Australian Journal of Agricultural and Resources Economics, 53, 437-456. Cooper, T.F., G. De 'ath, K.E. Fabricius, 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, 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., 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. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cottrell, B., C. Insley, R. Meade, 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, N. Tapper, 2010: Changing Urban Climate and CO2 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., 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, 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. Crook, J.A., L.A. Jones, P.M. Forster, 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, 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, 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, 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, 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 ResearchAtmospheres, 115(D2), D02101. CSIRO, BoM, 2007: Climate change in Australia. CSIRO and Bureau of Meteorology, Melbourne, 140 pp. CSIRO, Maunsell Australia Pty Ltd, Phillips Fox, 2007: Infrastructure and climate change risk assessment for Victoria. Report prepared for the Victorian Government. CSIRO, Aspendale, 76 pp. CSIRO, 2008: Water availability in the Murray-Darling Basin. A report from CSIRO to the Australian Government. CSIRO, Collingwood, 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. CSIRO, Black Mountain, ACT, 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). CSIRO, Collingwood, 31 pp. CSIRO, 2011: CSIRO, submission No. 40 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. CSIRO, Canberra (online at www.pc.gov.au/projects/inquiry/climate-changeadaptation/submissions), 31 pp. 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). CSIRO, Melbourne, 41 pp. CSIRO, BoM, 2012: State of the Climate 2012. CSIRO, Canberra, 12 pp. Cullen, J.M., L.E. Chambers, P.C. Coutin, P. Dann, 2009: Predicting onset and success of breeding in little penguins Eudyptula minor from ocean temperatures. Marine Ecology-Progress Series, 378, 269-278. Cullen, B.R., 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(0), 721-735. DAFF, 2012: Australia's agriculture, fisheries and forestry at a glance 2012. Department of Agriculture, Fisheries and Forestry, Canberra, 140 pp. Dalton, S.J., S. Godwin, S.D.A. Smith, 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, E.W.R. Barlow, 2013: Evaluation of recent trends in Australian pome fruit spring phenology. International Journal of Biometeorology, 57(3), 409-421. Davies, P.M., 2010: Climate Change Implications for River Restoration in Global Biodiversity Hotspots. Restoration Ecology, 18(3), 261-268. Davies, J., D. Campbell, M. Campbell, J. Douglas, H. Hueneke, M. LaFlamme, D. Pearson, K. Preuss, J. Walker, 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. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Davis, J., A. Pavlova, R. Thompson, P. Sunnucks, 2013: Evolutionary refugia and ecological refuges: key concepts for conserving Austrlaian arid zone reshwater biodiversity under climate change. Global Change Biology, 19, 1970-1984. DCC, 2009: Climate change risks to Australia’s coast: A first-pass national assessment. Department of Climate Change, Government of Australia, Canberra, Australia, 172 pp. DCCEE, 2010: Developing a National Coastal Adaptation Agenda: A report on the national climate change forum, Adelaide, 18-19 February 2010. Department of Climate Change and Energy Efficiency, Canberra, 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, Canberra, 16 pp. De'ath, G., J.M. Lough, 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, 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., 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, 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. 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, S.K. Wilson, 2013: Bleaching, coral mortality adn subsequent survivorship on a West Australian fringing reef. Coral Reefs, 32, 233-238. DERM, DIP, 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, Department of Infradtructure and Planning, and Local Government Association of Queensland. State of Queensland, Brisbane, Qld, 16 pp. Derraik, J.G.B., 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, 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, 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, O. Hoegh-Guldberg, 2009: Doom and Boom on a Resilient Reef: Climate Change, Algal Overgrowth and Coral Recovery. Plos One, 4(4), e5239. DIICCSRT, 2013: Climate Adaptation Outlook. A Proposed National Adaptation Assessment Framework. Department of Industry, Innovation, Climate Change, Science, Research and Tertiary Education, Canberra, 121 pp. DNP, 2010: Kakadu National Park Climate Change Strategy 2010-2015. Director of National Parks, Department of the Environment, Water, Heritage and the Arts, Canberra, 15 pp. Dobes, L., 2010: Notes on applying ‘real options’ to climate change adaptation measures, with examples from Vietnam. Crawford School, Australian National University, Canberra, 25 pp. Doherty, T.J., S. Clayton, 2011: The Psychological Impacts of Global Climate Change. American Psychologist, 66(4), 265–276. Donohue, R.J., T.R. McVicar, 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., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, 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. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Watson, R.J. Williams, 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, 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. Department of Sustainability and Environment Victoria, Melbourne, 28 pp. DSE, 2011: Guidelines for the development of a water supply demand strategy. Department of Sustainability and Environment Victoria, Melbourne, 100 pp. DSE, 2013: Victorian Climate Change Adaptation Plan. Department of Sustainability and Environment, State Government of Victoria, Melbourne, 96 pp. DSEWPC, 2011: Regional Natural Resources Management Planning for Climate Change Fund (online at www.environment.gov.au/cleanenergyfuture/regional-fund/). Department of the Environment (formerly Department of Sustainability, Environment, Water, Population and Communities), Canberra. Duckett, P.E., P.D. Wilson, A.J. Stow, 2013: Keeping up with teh neighbours: using a genetic measurement of dispersal and species dsitribution modelling to assess impact of climate change on an Australian aid zone gecko (Gehyra variegata). Diversity and Distributions, 19, 964-976. Duncan, M., G. Smart, 2011: Tool 2.1.4: Inundation modelling of present day and future floods, in: Impacts of Climate Change on Urban Infrastructure & the Build Environment - A Toolbox. NIWA, Wellington, 13 pp. Dunlop, M., P.R. Brown, 2008: Implications of climate change for Australia’s National Reserve System: a preliminary assessment. Department of Climate Change, Canberra, 188 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, H. Murphy, 2012: The Implications of Climate Change for Biodiversity Conservation and the National Reserve System: Final Synthesis. A report prepared for the Department of Sustainability, Environment, Water, Population and Communities, and the Department of Climate Change and Energy Efficiency. CSIRO Climate Adaptation Flagship, Canberra, 80 pp. Dunn, M.R., R.J. Hurst, J.A. Renwick, R.I.C.C. Francis, J. Devine, A. McKenzie, 2009: Fish abundance and climate trends in New Zealand. New Zealand Aquatic Environment and Biodiversity Report No. 31. Ministry of Fisheries, Wellington, 75 pp. Dupont, A., G. Pearman, 2006: Heating up the planet: climate change and security. Lowy Institute for International Policy, Sydney, 143 pp. Dupont, A., 2008: The Strategic Implications of Climate Change. Survival: Global Politics and Strategy, 50(3), 2954. Dwyer, A., C. Zoppou, O. Nielsen, S. Day, S. Roberts, 2004: Quantifying social vulnerability: a methodology for identifying those at risk to natural hazards. Geoscience Australia, Canberra, 92 pp. ECAN, 2005: Regional Coastal Environment Plan for the Canterbury Region. Environment Canterbury, Christchurch, 279 pp. Eckard, R.J., 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(0), 736-748. Edge Environment, 2011: Edge Environment, submission No. 54 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. Productivity Commission, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions), 7 pp. Edwards, B., 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, S. Friel, G. Hall, K. Larsen, S. Lockie, B. Wood, M. Lawrence, I. Hanigan, A. Hogan, L. Hattersley, 2011: Climate Change Adaptation at the Intersection of Food and Health. Asia-Pacific Journal of Public Health, 23(2, Supplement), 91-104. Elith, J., M. Kearney, 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. CSIRO, Canberra, 120 pp. Evans, L.S., P. Fidelman, C. Hicks, C. Morgan, A.L. Perry, R. Tobin, 2011: Limits to climate change adaptation in the Great Barrier Reef: scoping ecological and social limits. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 72 pp. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Evans, J.P., I. Boyer-Souchet, 2012: Local sea surface temperatures add to extreme precipitation in northeast Australia during La Niña. Geophys. Res. Lett., 39(10), L10803. Fabricius, K.E., C. Langdon, S. Uthicke, C. Humphrey, S. Noonan, G. De'ath, R. Okazaki, N. Muehllehner, M.S. Glas, 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., 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, 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, Melbourne, 54 pp. Feng, M., M.J. McPhaden, S.P. Xie, J. Hafner, 2013: La Nina forces unprecedented Leeuwin Current warming in 2011. Scientific Reports, 3, 1277. Fensham, R.J., R.J. Fairfax, D.P. Ward, 2009: Drought-induced tree death in savanna. Global Change Biology, 15(2), 380-387. Fidelman, P.I.J., A.M. Leitch, 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., P. Biro, D.J. Booth, 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. Figueira, W.F., D.J. Booth, 2010: Increasing ocean temperatures allow tropical fishes to survive overwinter in temperate waters. Global Change Biology, 16(2), 506-516. Finger, R., P. Lazzarotto, 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, 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: Efficient2009: 5th IWA Specialist Conference on Efficient Use and Management of Urban Water Supply, 25-28 October 2009, International Water Association, Sydney, 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, P. Trebicki, 2010: Future effects of elevated CO2 on wheat production – an overview of FACE research in Victoria, Australia. In: Food Security from Sustainable Agriculture. 15th Agronomy Conference, 15-18 November 2010, Lincoln, New Zealand [Dove, H., Culvenor, R.A. (eds.)], Australian Society of Agronomy, Gosford, NSW. Fitzharris, B.B., 2004: Possible impact of future climate change on seasonal snow of the Southern Alps of New Zealand. In: A Gaian World: Essays in Honour of Peter Holland [Fitzharris, G.K.a.B. (eds.)]. Department of Geography, University of Otago, Dunedin, pp. 231-241 Fitzharris, B.B., 2010: Climate change impacts on Dunedin. Dunedin City Council, Dunedin, 58 pp. Fitzpatrick, M.C., A.D. Gove, N.J. Sanders, 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, 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, 41 pp. Fletcher, C.S., B.M. Taylor, A.N. Rambaldi, B.P. Harman, S. Heyenga, K.R. Ganegodage, F. Lipkin, R.R.J. McAllister, 2013: Costs and coasts: an empirical assessment of physical and institutional climate adaptation pathways. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, 53 pp. FOA, MPI, 2012: New Zealand Plantation Forestry Industry: Facts and Figures 2011/2012. New Zealand Forest Owners Association Inc. and Ministry for Primary Industries, Wellington, 46 pp. Fouquet, A., G.F. Ficetola, A. Haigh, 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, K. Delaney, 2007: Work in progress: four future scenarios for New Zealand. Manaaki Whenua Press, Lincoln, New Zealand, 112 pp. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Frederiksen, J.S., C.S. Frederiksen, 2007: Interdecadal changes in southern hemisphere winter storm track modes. Tellus Series a-Dynamic Meteorology and Oceanography, 59(5), 599-617. Frederiksen, C.S., J.S. Frederiksen, J.M. Sisson, S.L. Osbrough, 2011: Changes and projections in Australian winter rainfall and circulation: Anthropogenic forcing and internal variablilty. Int. J. Climate Change Impacts and Responses, 2, 143-162. Freeman, C., 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, 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. Fuentes, M.M.P.B., M. Hamann, V. Lukoschek, 2009: Marine Reptiles. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 18 pp. 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, 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 the Global Forum 2011 [Otto-Zimmermann, K. (eds.)]. Springer, Dordrecht, the Netherlands, pp. 283-294. Funk, J., 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. Englert Duursma, J. O'Donnell, P.D. Wilson, P.O. Downey, L. Hughes, 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, advance online, DOI 10.1007/s10530-012-0342-6. Gallagher, R.V., L. Hughes, M.L. Leishman, 2012b: Species loss and gain in communities under future climate change: consequences for functional diversity Ecography, 36, 531-540. Gallant, A.J.E., K.J. Hennessy, J. Risbey, 2007: Trends in rainfall indices for six Australian regions: 1910-2005. Australian Meteorological Magazine, 56, 223-239. Gallant, A.J.E., D.J. Karoly, 2010: A Combined Climate Extremes Index for the Australian Region. Journal of Climate, 23, 6153-6165. Gallant, A.J.E., A.S. Kiem, D.C. Verdon-Kidd, R.C. Stone, 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, A. Clark, 2009: Climate Change Effects on the Land Transport Network Volume One: Literature Review and Gap Analysis. Report No 378. NZ Transport Agency, Wellington, New Zealand, 226 pp. Gardner, J., A.-M. Dowd, C. Mason, P. Ashworth, 2009a: A framework for stakeholder engagement on climate adaptation. CSIRO Climate Adaptation Flagship Working paper No. 3. CSIRO, Kenmore, Qld, 32 pp. Gardner, J.L., R. Heinsohn, L. Joseph, 2009b: Shifting latitudinal clines in avian body size correlate with global warming in Australian passerines. Proceedings of the Royal Society B-Biological Sciences, 276(1674), 38453852. Gardner, J., R. Parsons, G. Paxton, 2010: Adaptation benchmarking survey: initial report. CSIRO Climate Adaptation Flagship Working paper No. 4. CSIRO, Canberra, 58 pp. Garnaut, R., 2008: The Garnaut Climate Change Review: Final Report. Cambridge University Press, Melbourne, 634 pp. Garnett, S., D. Franklin, G. Ehmke, J. VanDerWal, L. Hodgson, C. Pavey, A. Reside, J. Welbergen, S. Butchart, G. Perkins, S. Williams, 2013: Climate change adaptation strategies for Australian birds. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 109 pp. Gaydon, R.S., H.G. Beecher, R. Reinke, S. Crimp, S.M. Howden, 2010: Rice. In: Adapting Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 67-83. GBRMPA, 2007: Great Barrier Reef Climate Change Action Plan 2007-2012. Great Barrier Reef Marine Park Authority, Commonwealth of Australia, Townsville, Qld, 14 pp. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 GBRMPA, 2009a: Great Barrier Reef Outlook Report 2009. Great Barrier Reef Marine Park Authority, Townsville, Qld, 212 pp. GBRMPA, 2009b: Great Barrier Reef Tourism Climate Change Action Strategy 2009-2012. Great Barrier Reef Marine Park Authority, Townsville, Qld, 56 pp. GBRMPA, 2011: Great Barrier Reef Report Card 2011: Reef Water Quality Protection Plan. Queensland Government, Brisbane, Qld, 6 pp. Gerard, P.J., J.R.F. Barringer, J.G. Charles, S.V. Fowler, J.M. Kean, C.B. Phillips, A.B. Tait, G.P. Walker, 2012: Potential effects of climate change on biological control systems: case studies from New Zealand. BioControl, advance online, DOI: 10.1007/s10526-012-9480-0. GETF, 2011: Roadmap to renewable and low emission energy in remote communities. Report by the Green Energy Task Force. Territory Climate Change, Northern Territory Government, Darwin, 65 pp. Gibbs, M., T. Hill, 2012: Coastal Climate Change Risk - Legal and Policy Responses in Australia. Department of Climate Change and Energy Efficiency, Canberra, 89 pp. Gibson, L., A. McNeill, P. de Tores, A. Wayne, 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., 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, D. Uzzell, 2009: Temporal pessimism and spatial optimism in environmental assessments: An 18-nation study. Journal of Environmental Psychology, 29(1), 1-12. Gifford, R., 2011: The Dragons of Inaction. Psychological Barriers That Limit Climate Change Mitigation and Adaptation. American Psychologist, 66(4), 290-302. Gillanders, B.M., T.S. Elsdon, I.A. Halliday, G.P. Jenkins, J.B. Robins, 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, 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., Jack, M. (eds.)]. Scion, Rorotua, NZ, pp. 71-121. Glavovic, B., W. Saunders, J. Becker, 2010: Land-use planning for natural hazards in New Zealand: the setting, barriers, ‘burning issues’ and priority actions. Natural Hazards, 54(3), 679-706. Gorddard, R., R. Wise, K. Alexander, A. Langston, A. Leitch, M. Dunlop, A. Ryan, J. Langridge, 2012: Striking the balance: coastal development and ecosystem values. Department of Climate Change and Energy Efficiency, Canberra, 114 pp. Gordon, D.P., J. Beaumont, A. MacDiarmid, D.A. Robertson, S.T. Ahyong, 2010: Marine Biodiversity of Aotearoa New Zealand. Plos One, 5(8), e10905. Gorman-Murray, A., 2008: Before and after Climate Change: The Snow Country in Australian Imaginaries. M/C Journal, 11(5), online. 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, S.N., G.R. McGregor, 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. Gosling, E., 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. Goss, J., B. Lindquist, 2000: Placing movers: An overview of the Asia-Oacific migration system. Contemporary Pacific, 12(2), 385-414. Grace, W.J., V.O. Sadras, P.T. Hayman, 2009: Modelling heatwaves in viticultural regions of southeastern Australia. Australian Meteorological and Oceanographic Journal, 58(4), 249-262. Grafton, R.Q., K. Hussey, 2007: Buying back the living Murray: at what price? Australasian Journal of Environmental Management, 14(2), 74-81. Graham, P., L. Reedman, F. Poldy, 2008: Modelling of the future of transport fuels in Australia. IR 1046. CSIRO, Canberra, 112 pp. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Gray, M., J.C. Altman, N. Halasz, 2005a: 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, The Australian National University, Canberra, 40 pp. Gray, W., R. Ibbitt, R. Turner, M. Duncan, M. Hollis, 2005b: A Methodology to assess the impacts of climate change on flood risk in New Zealand. NIWA Client Report CHC2005-060. Ministry for the Environment, Wellington, 40 pp. Green, A.E., M.R. Power, D.M. Jang, 2008a: Trans-Tasman migration: New Zealanders’ explanations for their move. New Zealand Geographer, 64(1), 34-45. Green, K., J.A. Stein, M.M. Driessen, 2008b: 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. Green, D., S. Jackson, J. Morrison, 2009: Risks from climate change to Indigenous communities in the tropical north of Australia. Department of Climate Change and Energy Efficiency, Canberra, 185 pp. Green, B.S., C. Gardner, A. Linnane, P.J. Hawthorne, 2010a: The Good, the Bad and the Recovery in an Assisted Migration. Plos One, 5(11), e14160. Green, D., L. Alexander, K. McLnnes, J. Church, N. Nicholls, N. White, 2010b: 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, A. Tapim, 2010c: Indigenous Australians' knowledge of weather and climate. Climatic Change, 100(2), 337-354. 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, 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), 13011330. Griffiths, G.M., 2007: Changes in New Zealand daily rainfall extremes 1930–2004. Weather and Climate, 27, 3-44. Guillaume, J., G.M. Li, M.F. Hutchinson, K. Proust, S. Dovers, 2010: Integrated Assessment of Climate Change Impacts on Urban Settlements (IACCIUS) Project: A differential vulnerability assessment of climate change impacts on Darwin. Fenner School of Environment and Society, The Australian National University, Canberra. Gunasekera, D., Y. Kim, C. Tulloh, 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, 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., C. Squires, 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, 130 pp. Gurran, N., 2008: The Turning Tide: Amenity Migration in Coastal Australia. International Planning Studies, 13(4), 391-414. Gurran, N., E. Hamin, 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, 66 pp. Hagger, V., D. Fisher, S. Schmidt, 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., 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., 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., K. Haynes, 2008: Community Bushfire Safety. CSIRO Publishing, Collingwood, Vic, 208 pp. Hanigan, I.C., C.D. Butler, P.N. Kokic, M.F. Hutchinson, 2012: Suicide and drought in New South Wales, Australia, 1970-2007. Proceedings of the National Academy of Sciences, 109(35), 13950-13955. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hanna, E.G., T. Kjellstrom, C. Bennett, K. Dear, 2011: Climate change and rising heat: population health implications for working people in Australia. Asia Pacific Journal of Public Health, 23(Supplement 2), S1426. Hannah, J., 2004: An updated analysis of long-term sea level change in New Zealand. Geophysical Research Letters, 31(3), L03307. Hannah, J., R. Bell, 2012: Regional sea level trends in New Zealand. Journal of Geophysical Research-Oceans, 117, C01004. Hansen, A., P. Bi, M. Nitschke, P. Ryan, D. Pisaniello, 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, D. Newth, 2013: Economic impacts of climate change on the Australian dairy sector. Australian Journal of Agricultural and Resource Economics, advance online, DOI: 10.1111/14678489.12021. Harley, D., P. Bi, G. Hall, A. Swaminathan, S.L. Tong, 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, Supplement), 54-66. Harman, I.N., M. Ford, G. Jakeman, S.J. Phipps, M. Brede, J.J. Finnigan, D. Gunasekera, H. Ahammad, 2008: Assessment of future global scenarios for the Garnaut Climate Change Review: an application of the GIAM framework. CSIRO, Canberra, 64 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, 33 pp. Harmsworth, G., B. Raynor, 2005: Cultural Consideration in Landslide Risk Perception. In: Landslide Hazard and Risk [Glade, T., Anderson, M.G., Crozier, M.J. (eds.)]. John Wiley & Sons, Ltd, London, pp. 219-249. Harmsworth, G., M. Tahi, C.K. Insley, 2010: Climate change business opportunities for Māori land and Māori organisations: Sustainable Land Management Mitigation and Adaptation to Climate Change (SLMACC). Manaaki Whenua Press, Palmerston North, 59 pp. Harper, B., T. Hardy, L. Mason, R. Fryar, 2009: Developments in storm tide modelling and risk assessment in the Australian region. Natural Hazards, 51(1), 225-238. Harris, R.J., G. Barker, 2007: Relative risk of invasive ants (Hymenoptera: Formicidae) establishing in New Zealand. New Zealand Journal of Zoology, 34(3), 161-178. 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, B.W. Brook, 2012: Managing the long-term persistence of a rare cockatoo under climate change. Journal of Applied Ecology, 49(4), 785-794. Harsch, M.A., R. Buxton, R.P. Duncan, P.E. Hulme, P. Wardle, 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, 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., K.J.E. Walsh, 2008: Tropical cyclone trends in the Australian region. Geochemistry Geophysics Geosystems, 9(7), Q07V07. Hasson, A.E.A., G.A. Mills, B. Timbal, 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, P.K. Campbell, 2011: Projections of the future costs of electricity generation technologies. EP104982. CSIRO, Canberra, 79 pp. HCCREMS, 2010: Potential Impacts of Climate Change on the Hunter, Central and Lower North Coast of NSW. Hunter Councils, Thornton, NSW, 86 pp. HCCREMS, 2012: Decision Support for Coastal Adaptation: The Handbook. Hunter Councils, Thornton, NSW, 178 pp. HDC, 2012: Long Term Plan 2012/22. Hastings District Council, Hastings, 80 pp. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Heckbert, S., J. Russell-Smith, A. Reeson, J. Davies, G. James, 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, J.S. Dukes, 2008: Five Potential Consequences of Climate Change for Invasive Species. Conservation Biology, 22(3), 534-543. Hendon, H.H., D.W.J. Thompson, 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., 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), 619630. Hendrikx, J., E. Hreinsson, M. Clark, 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, 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, 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., Canziani, O., Palutikof, J., van der Linden, P., Hanson, C. (eds.)]. Cambridge University Press, Cambridge, UK, 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, N. Plummer, 2008a: An assessment of the impact of climate change on the nature and frequency of exceptional climatic events. CSIRO and Bureau of Meteorology, Melbourne and Canberra, 33 pp. Hennessy, K.J., P.H. Whetton, K. Walsh, I.N. Smith, J.M. Bathols, M. Hutchinson, 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, M.C. Ellul, 2013: Enhancing climate change communication: strategies for profiling and targeting Australian interpretive communities. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 95 pp. Hitchings, T.R., 2009: Leptophlebiidae (Ephemeroptera) of the alpine region of the Southern Alps, New Zealand. Aquatic Insects, 31, 595-601. Hobday, A.J., E.S. Poloczanska, R. Matear (eds.), 2008: Implications of Climate Change for Australian Fisheries and Aquaculture: A preliminary assessment. Department of Climate Change, Canberra, 86 pp. 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., E.S. Poloczanska, 2010: Fisheries and Aquaculture. In: Adapting Agriculture to Climate Change: Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 205-228. Hobday, A.J., 2011: Sliding baselines and shuffling species: implications of climate change for marine conservation. Marine Ecology, 32(3), 392-403. Hobson, K., 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, B. Loechel, 2010a: Climate adaptation in the Australian mining and exploration industries. CSIRO Climate Adaptation Flagship Working paper No. 5. CSIRO, Canberra, 32 pp. Hodgkinson, J.H., B. Loechel, K. Moffat, M. Howden, A. Littleboy, S. Crimp, 2010b: Climate risks to the Australian mining industry - a preliminary review of vunerabilities. In: Sustainable Mining 2010 Conference: 17-19 August 2010, Kalgoorlie, Western Australia. The Australasian Institute of Mining and Metallurgy, Carlton South, Vic, pp. 341-350. 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, M.E. Hatziolos, 2007: Coral reefs under rapid climate change and ocean acidification. Science, 318(5857), 1737-1742. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hoegh-Guldberg, O., 2012: The adaptation of coral reefs to climate change: is the Red Queen being outpaced? Scientia Marina, 76(2), 403-408. Hoelzle, M., T. Chinn, D. Stumm, F. Paul, M. Zemp, 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, 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, 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, 64 pp. Hogan, A., A. Bode, 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, J.S. Risbey, 2009: El NiñoSouthern Oscillation. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 25 pp. Hope, P., W. Drosdowsky, N. Nicholls, 2006: Shifts in the synoptic systems influencing southwest Western Australia. Climate Dynamics, 26(7-8), 751-764. Hope, P., B. Timbal, 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, S. Becken, 2012: Climate change in a regional context: relative vulnerability in the Australasian skier market. Regional Environmental Change, advance online, DOI 10.1007/s10113-012-0352-z. Hovenden, M.J., A.L. Williams, 2010: The impacts of rising CO2 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, J. Williamson, 2012: Ocean acidification. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2012 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 38 pp. Howden, S.M., 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. (eds.)]. CSIRO, Collingwood, Vic, pp. 227-244. Howden, S.M., S.J. Crimp, 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, R. Nelson, 2010: Australian agriculture in a climate of change. In: Managing Climate Change: Papers from the GREENHOUSE 2009 Conference [Jubb, I., Holper, P., Cai, W. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 101-111. Howden, S.M., C.J. Stokes, 2010: Introduction. In: Adapting Agriculture to Climate Change: Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 1-11. 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., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 112-121. Howe, C., R.N. Jones, S. Maheepala, B. Rhodes, 2005: Melbourne Water Climate Change Study: Implications of Potential Climate Change for Melbourne’s Water Resources. Report CMIT-2005-104. CSIRO and Melbourne Water, Melbourne, 26 pp. Howells, E.J., R. Berkelmans, M.J.H. van Oppen, B.L. Willis, L.K. Bay, 2013: Historical thermal regimes define limits to coral acclimatization. Ecology, 94(5), 1078-1088. HRC, 2012: Proposed One Plan. Chapter 10: Natural Hazards. Horizons Regional Council, Palmerston North, 8 pp. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Huang, C., A.G. Barnett, X. Wang, S. Tong, 2012: The impact of temperature on years of life lost in Brisbane, Australia. Nature Clim. Change, 2(4), 265-270. Hughes, L., 2003: Climate change and Australia: trends, projections and impacts. Austral Ecology, 28, 423-443. Hughes, T.P., M.J. Rodrigues, D.R. Bellwood, D. Ceccarelli, O. Hoegh-Guldberg, L. McCook, N. Moltschaniwskyj, M.S. Pratchett, R.S. Steneck, B. Willis, 2007: Phase Shifts, Herbivory, and the Resilience of Coral Reefs to Climate Change. Current Biology, 17(4), 360-365. Hughes, R., D. Mercer, 2009: Planning to Reduce Risk: The Wildfire Management Overlay in Victoria, Australia. Geographical Research, 47(2), 124-141. Hughes, L., R. Hobbs, A. Hopkins, J. McDonald, M. Stafford-Smith, W. Steffen, S.E. Williams, 2010: National Climate Change Adaptation Research Plan: Terrestrial Biodiversity. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 70 pp. Hughes, J.D., K.C. Petrone, R.P. Silberstein, 2012: Drought, groundwater storage and streamflow decline in southwestern Australia. Geophysical Research Letters, 39, L03408. Hugo, G., 2010: Climate change-induced mobility and the existing migration regime in Asia and the Pacific. In: Climate Change and Displacement [McAdam, J. (eds.)]. Hart Publishing, Oxford, 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. National Adaptation Research Facility (NCCARF), Gold Coast, Qld, 101 pp. Hunt, J., J. Altman, K. May, 2009: Social benefits of Aboriginal engagement in natural resource management. CAEPR Working Paper 60. Centre for Aboriginal Economic Policy Research, The Australian National University, Canberra, 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., 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., S. Dovers (eds.), 2007: Managing water for Australia: the social and institutional challenges. CSIRO Publishing, Collingwood, Vic, 157 pp. IAA, 2011a: Institute of Actuaries of Australia, submission No. 43 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. Productivity Commission, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions), 60 pp. IAA, 2011b: Institute of Actuaries of Australia, submission to the Garnaut Climate Change Review - Update 2011. Garnaut Climate Change Review, Canberra, 28 pp. IAG, 2011: Insurance Australia Group, submission No. 39 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. Productivity Commission, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions), 24 pp. ICA, 2012: Historical Disaster Statistics (online database, accessed 31 August 2013, at http://www.insurancecouncil.com.au/industry-statistics-data/disaster-statistics/historical-disaster-statistics). Insurance Council of Australia, Sydney, NSW. ICNZ, 2013: The Cost of Disaster events (online database, accessed 31 August 2013, at www.icnz.org.nz/naturaldisaster/historic-events). Insurance Council of New Zealand, Wellington. 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., R. Meade, 2008: Māori Impacts from the Emissions Trading Scheme: Detailed Analysis and Conclusions. Ministry for the Environment, Wellington, New Zealand, 57 pp. Insley, C.K., 2010: Survey of Māori business: climate change Māori business opportunities. 37 Degrees South Aotearoa, Gisborne, New Zealand, 35 pp. IOCI, 2012: Indian Ocean Climate Initiative Stage 3: Summary for Policymakers. CSIRO and BOM, Melbourne, 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, 582 pp. Irving, A.D., S.D. Connell, B.D. Russell, 2011: Restoring Coastal Plants to Improve Global Carbon Storage: Reaping What We Sow. Plos One, 6(3), e18311. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Irving, D., P. Whetton, 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, G. Kuczera, 2013: Evaluating the non-stationarity of Australian annual maximum flood. Journal of Hydrology, 494, 134–145. Jackson, A.C., M.G. Chapman, 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., L. Kelly, 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. Jakes, P.J., E.R. Langer, 2012: The adaptive capacity of New Zealand communities to wildfire. International Journal of Wildland Fire, 21(6), 764-772. Jakob, D., D. Karoly, 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, 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, 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, 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, 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, R.K. Scott, 2002: Exposure to bushfire smoke and asthma: an ecological study. Medical Journal of Australia, 176, 535-538. Johnston, G., D. Burton, M. Baker-Jones, 2013: Climate Change Adaptation in the Boardroom: combined report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 85 pp. Jollands, N., M. Ruth, C. Bernier, 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, R.N., F. Chiew, W. Boughton, 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. Jones, D., W. Wang, R. Fawcett, 2009: High-quality spatial climate data-sets for Australia. Australian Meteorological and Oceanographic Journal, 58(4), 233-248. Jones, R., A. Wardell-Johnson, M. Gibberd, A. Pilgrim, G. Wardell-Johnson, J. Galbreath, S. Bizjak, D. Ward, K. Benjamin, J. Carlsen, 2010: The impact of climate change on the Margaret River Wine Region: developing adaptation and response strategies for the tourism industry. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 94 pp. Kamber, G., C. McDonald, G. Price, 2013: Drying out: Investigating the economic effects of drought in New Zealand. Reserve Bank of New Zealand Analytical Note series AN2013/02. Reserve Bank of New Zealand, Wellington, 31 pp. Kamman, C., L. Grünhage, U. Grüters, S. Janze, H.-J. Jäger, 2005: Response of aboveground grassland biomass and soil moisture to moderate long-term CO2 enrichment. Basic and Applied Ecology, 6, 351-365. Karoly, D.J., K. Braganza, 2005: Attribution of recent temperature changes in the Australian region. Journal of Climate, 18, 457-464. Karoly, D.J., 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., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 412-418. Kawaguchi, S., A. Ishida, R. King, B. Raymond, N. Waller, A. Constable, S. Nicol, M. Wakita, A. Ishimatsu, 2013: Risk maps for Antarctic krill under projected Southern Ocean acidification. Nature Climate Change, advance online, DOI: 10.1038/nclimate1937. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kay, R., A. Travers, L. Dalton, 2014: The evolution of Coastal Vulnerability Assessments to Support Adaptive Decision Making in Australia: A Review. In: Climate Change and the Coastal Zone: Building Resilient Communities [Glavovic, B., Kay, R.C., Kelly, M., Travers, A. (eds.)]. CRC Press, London, www.crcpress.com/product/isbn/9780415464871. KCDC, 2012: Proposed Kapiti Coast District Plan, Volume 1. Notified 29 November 2012. Kapiti Coast District Council, Paraparaumu. Kearney, M.R., W.P. Porter, C. Williams, S. Ritchie, 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, 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, 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, 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, M. Bedward, 2010: Decadal change in wetland-woodland boundaries during the late 20th 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, 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., W. Zhang, 2009: Allocation of New Zealand Units within Agriculture in the New Zealand Emissions Trading System. Working Paper 09-16. Motu Economic and Public Policy Research, Wellington, 82 pp. Khalaj, B., G. Lloyd, V. Sheppeard, 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, M. Wehner, 2013: Changes in temperature and precipitation extremes in the CMIP5 ensemble. Climatic Change, 119(2), 345-357. Kiem, A.S., S.W. Franks, G. Kuczera, 2003: Multi-decadal variability of flood risk. Geophysical Research Letters, 30(2), L015992. Kiem, A.S., D.C. Verdon-Kidd, S. Boulter, J. Palutikof, 2010: Learning from experience: Historical Case Studies and Climate Change Adaptation, Case studies of extreme events, Synthesis Report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 42 pp. Kiem, A.S., E.K. Austin, 2012: Limits and barriers to climate change adaptation for small inland communities affected by drought. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 46 pp. Kiem, A.S., 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, advance online, DOI: 10.1016/j.gloenvcha.2013.06.003. King, D., W. Tawhai, A. Skipper, W. Iti, 2005: Anticipating local weather and climate outcomes using Māori environmental indicators. NIWA, Auckland, New Zealand, 18 pp. King, D., G. Penny, 2006: The 2nd Māori Climate Forum - Hongoeka Marae, Plimmerton (24 May 2006): Summary Report. National Institute of Water & Atmospheric Research Ltd, Auckland, New Zealand, 26 pp. King, D., J. Goff, 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, D. Hosking, 2008: Ground-truthing pre-event recovery planning issues with Ngāti Rongomai. NIWA, Auckland, New Zealand, 72 pp. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 King, D., G. Penny, 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., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 100-111. King, K.J., R.M. de Ligt, 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. King, D., W. Dalton, M. Home, M. Duncan, M.S. Srinivasan, J. Bind, C. Zammit, A. McKerchar, D. AshfordHosking, 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. Report AKL2011-015, prepared for Te Rūnanga o Arowhenua Society Incorporated and the New Zealand Climate Change Research Institute, Victoria University. NIWA, Auckland, 133 pp. 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, K.A. Wilson, 2009: Major Conservation Policy Issues for Biodiversity in Oceania. Conservation Biology, 23(4), 834-840. Kingsford, R.T. (ed.), 2011: Conservation Management of Rivers and Wetlands under Climate Change, Marine & Freshwater Research, Special Issue Volume 62 Number 3. CSIRO Publishing, Collingwood, Vic, 217-327 pp. Kingsford, R.T., J.E.M. Watson, 2011: Climate change in Oceania: A synthesis of biodiversity impacts and adaptations. Pacific Conservation Biology, 17, 270-284. 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, S. Keyworth, 2012: The economic impact of water reductions during the Millennium Drought in the Murray-Darling Basin. In: 56th AARES Annual Conference, 7-10 February 2012, Fremantle, Australia, Australian Agricultural & Resource Economics Society, Canberra, 25 pp. Kirby, M., F. Chiew, M. Mainuddin, B. Young, G. Podger, A. Close, 2013: Drought and climate change in the Murray-Darling Basin: a hydrological perspective. In: Drought in Arid and Semi-Arid Environments: A MultiDisciplinary and Cross-Country Perspective [Schwabe, K., Albiac, J., Connor, J., Hassan, R., Meza-Gonzales, L. (eds.)]. Springer, Dordrecht, Netherlands, pp. 281-299. Kirono, D.G.C., D. Kent, 2010: Assessment of rainfall and potential evaporation from global climate modelsand its implications for Australian regional drought projection. International Journal of CLimatology, 31(9), 12951308. Kirono, D.G.C., D.M. Kent, K.J. Hennessy, 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., 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, 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. Minstry of Agriculture and Forestry, Wellington, 129 pp. Kirschbaum, M.U.F., D. Whitehead, S.M. Dean, P.N. Beets, J.D. Shepherd, A.G.E. Ausseil, 2011b: Implications of albedo changes following afforestation on the benefits of forests as carbon sinks. Biogeosciences, 8(12), 36873696. Kirschbaum, M.U.F., M.S. Watt, A. Tait, 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, 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, G. Williams, 2010: Impacts of Climate Change on Rural Water Infrastructure. Report to Ministry of Agriculture and Forestry. MWH New Zealand, Wellington, 122 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., L. Dwyer, 2010: Seasonal variation versus climate variation for Australian Tourism. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 27 pp. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kuleshov, Y., R. Fawcett, L. Qi, B. Trewin, D. Jones, J. McBride, H. Ramsay, 2010: Trends in tropical cyclones in the South Indian Ocean and the South Pacific Ocean. Journal of Geophysical Research-Atmospheres, 115, D01101. Kuruppu, N., J. Murta, P. Mukheibir, J. Chong, T. Brennan, 2013: Understanding the adaptive capacity of Australian small-to-medium enterprises to climate change and variability. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 133 pp. Kwan, D., H. Marsh, 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. Langton, M., M. Parsons, S. Leonard, K. Auty, D. Bell, C.P. Burgess, S. Edwards, R. Howitt, S. Jackson, V. McGrath, J. Morrison, 2012: National Climate Change Adaptation Research Plan for Indigenous Communities. NCCARF Publication 11/12. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 48 pp. Last, P.R., W.T. White, D.C. Gledhill, A.J. Hobday, R. Brown, G.J. Edgar, 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, C. Cocklin, 2011: The 10 Australian ecosystems most vulnerable to tipping points. Biological Conservation, 144(5), 1472-1480. Lawrence, J., S. Allan, 2009: A strategic framework and practical options for integrating flood risk management to reduce flood risk and the effects of climate change. Ministry for the Environment, Wellington, 79 pp. Lawrence, A., T. Baker, C.E. Lovelock, 2012: Optimising and managing coastal carbon: Comparative sequestration and mitigation opportunities across Australia’s landscapes and land uses. FRDC Report 2011/084. Fisheries Research and Development Corporation, Deakin, ACT, 66 pp. Lawrence, J., A. Reisinger, B. Mullan, 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, L. McGlinchey, 2013b: Adapting to changing climate risk by local government in New Zealand: institutional practice barriers and enablers. Local Environment, advance online, DOI: 10.1080/13549839.2013.839643. Leblanc, M., S. Tweed, A. Van Dijk, 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., 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., Lockie, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 117-131. Lennox, J., W. Proctor, 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, M. Stafford-Smith, 2013a: Understanding the role of compound events in climate extremes. WIRES, advance online, DOI: 10.1002/wcc.252. Leonard, S., J. Mackenzie, F. Kofod, M. Parsons, M. Langton, P. Russ, L. Ormond-Parker, K. Smith, M. Smith, 2013b: Indigenous climate change adaptation in the Kimberley region of North-western Australiat: Learning from the past, adapting in the future: Identifying pathways to successful adaptation in Indigenous communities. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 131 pp. Leslie, L., M. Leplastrier, 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, D. Brown, 2009: Farm dam interception in the Campaspe Basin under climate change. In: H2009: 32nd Hydrology and Water Resources Symposium, 30 November-3 December 2009, Newcastle, Australia, Engineers Australia, Barton, ACT. pp 1194-1204. Leviston, Z., A. Leitch, M. Greenhill, R. Leonard, I. Walker, 2011: Australians’ views of climate change. CSIRO, Canberra, 21 pp. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Leviston, Z., I. Walker, S. Morwinski, 2012: Your opinion on climate change might not be as common as you think. Nature Clim. Change, 3, 334–337. Lewandowsky, S., 2011: Popular consensus: climate change is set to continue. Psychological Science, 22(4), 460463. LGNZ, 2008: Local Government New Zealand, submission on the proposed Coastal Policy Statement. Department of Conservation, Wellington, 40 pp. Lieffering, M., P.C.D. Newton, F.Y. Li, 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., Nottage, R. (eds.)]. Ministry for Primary Industries, 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 CSIRO Publishing, Camberwell, Vic, 138 pp. 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, 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, 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), 719731. 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. Linnenluecke, M.K., A. Stathakis, 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, M. Howden, 2010: Goldfields-Esperance Regional Mining Climate Vulnerability Workshop: Report on workshop outcomes. CSIRO, Canberra, 42 pp. Loechel, B., J. Hodgkinson, K. Moffat, 2013: Climate Change Adaptation in Australian mining communities: comparing mining company and local government views and activities. Climatic Change, advance online, DOI: 10.1007/s10584-013-0721-8. Lough, J.M., 2008: Shifting climate zones for Australia's tropical marine ecosystems. Geophysical Research Letters, 35(14). Lough, J.M., 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, 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, 210(9), 41. Lovelock, C.E., G. Skilleter, N. Saintilan, 2009: Tidal Wetlands and Climate Change. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 32 pp. Low Choy, D., S. Serrao-Neumann, F. Crick, G. Schuch, M. Sanò, R.v. Staden, O. Sahin, B. Harman, S. Baum, 2012: Adaptation Options for Human Settlements in South East Queensland – Main Report. South East Queensland Climate Adaptation Research Initiative, Griffith University, Nathan, Qld, 210 pp. Lucas, C., K.J. Hennessy, J.M. Bathols, 2007: Bushfire weather in Southeast Australia: recent trends and projected climate change impacts. Bushfire Cooperative Research Centre, Melbourne, 84 pp. Lundquist, C.J., D. Ramsay, R. Bell, A. Swales, 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. McDonaldMadden, S.E. Williams, 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, 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, Gatton, Qld, 13 pp. Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 LWF, 2010: Report of the Land and Water Forum: A Fresh Start for Freshwater. Land and Water Forum, Wellington, 85 pp. Mac Nally, R., G. Horrocks, H. Lada, P.S. Lake, J.R. Thomson, 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, J. McDonald, 2013: Limp, leap or learn? Developing legal frameworks for climate change adaptation planning in Australia. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 277 pp. Mackey, B.G., J.E.M. Watson, G. Hope, 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, 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, R.A. Fuller, 2013: Protecting and restoring habitat to help Australia’s threatened species adapt to climate change. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 53 pp. Maloney, S.K., 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., 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, M. Waschka, 2010: National Climate Change Adaptation Research Plan for Marine Biodiversity and Resources. National Climate Change Adaptation Research Facility, Gold Coast, Qld, 68 pp. Marcar, N.E., R.G. Benyon, P.J. Polglase, K.I. Paul, S. Theiveyanathan, L. Zhang, 2006: Predicting the hydrological impacts of bushfire and climate change in forested catchments of the River Murray uplands: a review. CSIRO Water for a Healthy Country National Research Flagship, Canberra, 38 pp. Marshall, P., J. Johnson, 2007: The Great Barrier Reef and climate change: vulnerability and management implications. Chapter 24. In: Climate change and the Great Barrier Reef. A vulnerability assessment. [Johnson, J., Marshall, P. (eds.)]. Great Barrier Reef Marine Park Authority and Australian Greenhouse Office, Townsville, Australia, pp. 745-771. Marshall, N.A., 2010: Understanding social resilience to climate variability in primary enterprises and industries. Global Environmental Change, 20(1), 36-43. Maru, Y.T., J. Langridge, B.B. Lin, 2011: Current and potential applications of typologies in vulnerability assessments and adaptation science. CSIRO Climate Adaptation Flagship Working paper No. 7. CSIRO, Melbourne, 25 pp. Maru, Y.T., V. Chewings, 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, 91 pp. Mason, M., K. Haynes, 2010: Adaptation Lessons from Cyclone Tracy. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 82 pp. Mason, M., K. Haynes, G. Walker, 2013: Cyclone Tracy and the road to improving wind resistant design. In: Natural Disasters and Adaptation to Climate Change [Boulter, S., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 137-149. Matusik, G., K.X. Ruthrof, N.C. Brouwers, B. Dell, 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, CSIRO, 2008: Impacts of climate change on infrastructure in Australia and CGE model inputs. Report prepared by Maunsell Australia Pty Ltd in association with CSIRO Sustainable Ecosystems. Garnaut Climate Change Review, Canberra, 133 pp. Mayor, K., 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. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McAdam, J., 2010: ‘Disappearing states’, statelessness and the boundaries of international law. In: Climate Change and Displacement [McAdam, J. (eds.)]. Hart Publishing, Oxford, UK, pp. 105-130. McAlpine, C.A., J. Syktus, R.C. Deo, P.J. Lawrence, H.A. McGowan, I.G. Watterson, S.R. Phinn, 2007: Modeling the impact of historical land cover change on Australia's regional climate. Geophysical Research Letters, 34(22), L22711. McAlpine, C.A., J. Syktus, J.G. Ryan, R.C. Deo, G.M. McKeon, H.A. McGowan, 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., N. Nicholls, 1983: Seasonal relationships between Australian rainfall and the Southern Oscillation. Monthly Weather Review, 111(10), 1998-2004. 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, CSIRO, Bureau of Meteorology), Perth, 53 pp. McCallum, J.L., R.S. Crosbie, G.R. Walker, 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, 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., Macintosh, A., McDonald, J. (eds.)]. Federation Press, Annandale, NSW, 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., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 126-132. McGlone, M., S. Walker, R. Hay, 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., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 82-99. McGlone, M.S., 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 312. Department of Conservation, Wellington, 77 pp. 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, J.A. Bathols, 2005: Climate Change in Eastern Victoria. Stage 1 Report: The Effect of Climate Change on Coastal Wind and Weather Patterns. Gippsland Coastal Board, Melbourne, Vic, 26 pp. McInnes, K.L., I. Macadam, G.D. Hubbert, 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, 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, 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, 2012: Climate Futures for Tasmania: Extreme Sea Levels along Tasmania’s Coast and the Impact of Climate Change. Antarctic Climate and Ecosystems CRC, 44 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., 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., B.O. Wolf, 2010: Climate change increases the likelihood of catastrophic avian mortality events during extreme heat waves. Biology Letters, 6(2), 253-256. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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, K.A. Day, 2009: Climate change impacts on Australia's rangeland livestock carrying capacity: A review of challenges. Land & Water Australia, Canberra, 69 pp. McKerchar, A., B. Mullan, 2004: Seasonal inflow distributions for New Zealand hydroelectric power stations. NIWA Client Report: CHC2004-131. Ministry of Economic Development, Wellington, NZ, 12 pp. McKerchar, A.I., J.A. Renwick, 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, 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, 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, S. Tong, 2009: National climate change adaptation plan for human health. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 64 pp. McMillan, H., B. Jackson, S. Poyck, 2010: Flood Risk Under Climate Change. NIWA Report CHC2010-033. Ministry for Agriculture and Forestry, Wellington, 53 pp. McMillan, H., M. Duncan, G. Smart, J. Sturman, S. Poyck, S. Reese, A. Tait, E. Hreinsson, 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, W. R., K. Parnell, 2011: Limits to Climate Change Adaptation for Low-Lying Communities in the Torres Strait. Final report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 93 pp. McVicar, T.R., T.G. Van Niel, L.T. Li, M.L. Roderick, D.P. Rayner, L. Ricciardulli, 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. McVicar, T.R., R.J. Donohue, A.P. O'Grady, 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. CSIRO Water for a Healthy Country National Research Flagship. Murray-Darling Basin Authority, Canberra, 95 pp. MDBA, 2011: Proposed Basin Plan. Publication 192/11. Draft plan prepared for the Commonwealth of Australia. Murray-Darling Basin Authority, Canberra, 226 pp. MDBA, 2012: Proposed Basin Plan - A Revised Draft. Publication 57/12. Murray-Darling Basin Authority, Canberra, 235 pp. MDBC, 2007: Impact of the 2003 Alpine Bushfires on Streamflow: Broadscale Water Yield Assessment. Murray Darling Basin Commission, Canberra, 122 pp. Measham, T., B. Preston, T. Smith, C. Brooke, R. Gorddard, G. Withycombe, 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. Ministry of Economic Development, Wellington, 12 pp. Medlyn, B.E., R.A. Duursma, 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, 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), Gold Coast, Qld, 189 pp. 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, 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., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 747-845. Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Melbourne Water, 2010: Melbourne Water Responding to a Changing Climate. Melbourne Water, Melbourne, 19 pp. Melloy, P., G. Hollaway, J. Luck, R. Norton, E. Aitken, S. Chakraborty, 2010: Production and fitness of Fusarium pseudograminearum inoculum at elevated carbon dioxide in FACE. Global Change Biology, 16(12), 33633373. Memon, A., 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, E. Weber, 2010: Enhancing Potential for Integrated Catchment Management in New Zealand: A Multi-scalar, Strategic Perspective. Australasian Journal of Environmental Management, 17(1), 3544. Mendham, E., 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., 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. Meyssignac, B., A. Cazenave, 2012: Sea level: A review of present-day and recent-past changes and variability. 58(0), 96-109. MfE, 2007a: Environment New Zealand 2007. Report ME 847. Ministry for the Environment, Wellington, 456 pp. MfE, 2007b: Consultation with Māori on climate change: Hui Report. Report ME 830. Ministry for the Environment, Wellington, New Zealand, 135 pp. MfE, 2008a: Preparing for climate change: A guide for local government in New Zealand. Ministry for the Environment, Wellington, 44 pp. MfE, 2008b: Climate Change Effects and Impacts Assessment: A Guidance Manual for Local Government in New Zealand (2nd edition). Report ME870, prepared by B. Mullan, D. Wratt, S. Dean, M. Hollis, S. Allan, T. Williams, G. Kenny. Ministry for the Environment, Wellington, 167 pp. MfE, 2008c: Climate Change and Long-term Council Community Planning. Ministry for the Environment, Wellington, 22 pp. MfE, 2008d: Coastal Hazards and Climate Change: A Guidance Manual for Local Government in New Zealand (2nd edition). Revised by Ramsay, D., and Bell, R. (NIWA). Ministry for the Environment, Wellington, 139 pp. MfE, 2010a: Legally Protected Conservation Land in New Zealand: Environmental Snapshot April 2010. INFO 492. Ministry for the Environment, Wellington, 7 pp. MfE, 2010b: Tools for Estimating the Effects of Climate Change on Flood Flow: A Guidance Manual for Local Government in New Zealand. Report ME1013. New Zealand Ministry of Environment, Wellington, New Zealand, 63 pp. MfE, 2011: Freshwater Management 2011. National Policy Statement. Ministry for the Environment, Wellington, 12 pp. MfE, 2013: New Zealand’s Greenhouse Gas Inventory 1990–2011. Publication ME 1113. Ministry for the Environment, Wellington, 463 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, 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, K.J., A.A. Rowden, A. Williams, 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). Miller, G.M., S.A. Watson, J.M. Donelson, M.I. McCormick, P.L. Munday, 2012: Parental environment mediates impacts of increased carbon dioxide on a coral reef fish. Nature Climate Change, 2(12), 858-861. Mills, R., 2009: Way Ahead. Rio Tinto Review, March 2009. Rio Tinto, Melbourne, Vic, 2 pp. Min, S.-K., W. Cai, 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. Ministry of Conservation, Wellington, 31 pp. Mitchell, N.J., F.W. Allendorf, S.N. Keall, C.H. Daugherty, N.J. Nelson, 2010: Demographic effects of temperature-dependent sex determination: will tuatara survive global warming? Global Change Biology, 16(1), 60-72. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Moise, A.F., 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 ResearchAtmospheres, 113(D15), D009250. Moon, J.W., S.-H. Han, 2011: Thermostat strategies impact on energy consumption in residential buildings. Energy and Buildings, 43, 338-346. 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, 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. Moore, A.D., 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. Morrison, C., C.M. Pickering, 2011: Climate change adaptation in the Australian Alps: impacts, strategies, limits and management. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 84 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, T. Rayner, 2011: Climate change and its implications for Australia's freshwater fish. Marine and Freshwater Research, 62, 1082-1098. Mortimer, E., A. Bergin, R. Carter, 2011: Sharing risk: financing Australia’s disaster resilience. Australian Strategic Policy Institute, Barton, ACT, 24 pp. Mukheibir, P., N. Kuruppu, A. Gero, J. Herriman, 2013: Cross-Scale Barriers to Climate Change Adaptation in Local Government, Australia. Final report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 101 pp. Mulet-Marquis, S., J.R. Fairweather, 2008: Rural Population and Farm Labour Change. Lincoln University, Lincoln, New Zealand, 35 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, 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.)]. IGCI, Waikato, Hamilton, pp. 11-31. Mullan, B., A. Porteous, D. Wratt, M. Hollis, 2005: Changes in drought risk with climate change. NIWA Reort WLG2005-23. Ministry for the Environment, Wellington, 58 pp. Mullan, B., S.J. Stuart, M.G. Hadfield, M.J. Smith, 2010: Report on the Review of NIWA’s ‘Seven-Station’ Temperature Series. NIWA Information Series No. 78. National Institute of Water and Atmospheric Research (NIWA), Wellington, 175 pp. Mullan, B., T. Carey-Smith, G.M. Griffiths, A. Sood, 2011: Scenarios of storminess and regional wind extremes under climate change. NIWA client report WLG2010-31. Ministry of Agriculture and Forestry, Wellington, 80 pp. Munday, P.L., N.E. Crawley, 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, M. Dunlop, 2012: Queensland's biodiversity under climate change: terrestrial ecosystems. CSIRO Climate Adaptation Flagship Working Paper No. 12C. CSIRO, Canberra, 120 pp. Mustelin, J., N. Kuruppu, A.M. Kramer, J. Daron, K. de Bruin, A.G. Noriega, 2013: Climate Adaptation Research for the Next Generation. Climate and Development, advance online, DOI: 10.1080/17565529.2013.812953. Naish, S., K. Mengersen, W. Hu, S. Tong, 2013: Forecasting the futrue risk of Barmah Forest Virus disease under climate change sceanrios in Queensland, Australia. Plos One, 8(5), e62843. Nana, G., F. Stokes, W. Molano, 2011a: The Asset base, Income, Expenditure and GDP of the 2010 Māori Economy. Report for the Māori economic taskforce. Business and Economic Research Ltd, Wellington, New Zealand, 44 pp. Nana, G., F. Stokes, W. Molano, 2011b: The Māori economy, science and innovation scenarios. Report for the Māori economic taskforce. Business and Economic Research Ltd, Wellington, New Zealand, 42 pp. NDIR, 2011: Inquiry into flood insurance and related matters. Natural Disaster Insurance Review, Canberra, 149 pp. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nearing, M.A., F.F. Pruski, 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, 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, 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, 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, M. Lieffering, 2006: Impacts of elevated CO2 on a grassland grazed by sheep: the New Zealand FACE experiment. In: Managed Ecosystems and CO2: Case Studies, Processes, and Perspectives [Nosberger, J., Long, S.P., Norby, R.J., Stitt, M., Hendrey, G.R., Blum, H. (eds.)]. SpringerVerlag, 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, M.J. Hovenden, 2013: Selective grazing modifies previously anticipated responses of plant community composition to elevated CO2 in a temperate grassland. Global Change Biology, advance online, DOI: 10.1111/gcb.12301. NFRAG, 2008: National Flood Risk Advisory Group: Flood risk management in Australia. Australian Journal of Emergency Management, 23, 21-27. Nguyen, M., X. Wang, D. Chen, 2010: An investigation of extreme heatwave events and their effects on building and infrastructure. CSIRO, Canberra, 113 pp. Nicholls, N., 2006: Detecting and attributing Australian climate change: a review. Australian Meteorological Magazine, 55(3), 199-211. Nicholls, N., D. Collins, 2006: Observed climate change in Australia over the past century. Energy & Environment, 17(1), 1-12. 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., P. Uotila, L. Alexander, 2010: Synoptic influences on seasonal, interannual and decadal temperature variations in Melbourne, Australia. International Journal of Climatology, 30(9), 1372-1381. 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. Nidumolu, U., S. Crimp, D. Gobbett, A. Laing, M. Howden, S. Little, 2011: Heat stress in dairy cattle in northern Victoria: responses to a changing climate. Climate Adaptation Flagship Working paper No. 10. CSIRO, Canberra, 72 pp. Nilsson, G.E., D.L. Dixson, P. Domenici, M.I. McCormick, C. Sorensen, S.A. Watson, P.L. Munday, 2012: Nearfuture carbon dioxide levels alter fish behaviour by interfering with neurotransmitter function. Nature Climate Change, 2(3), 201-204. NIWA, MWH, GNS Science, BRANZ, 2012: Impacts of Climate Change on Urban Infrastructure and the Built Environment: Toolbox Handbook. National Institute of Water and Atmospheric Research (NIWA), Wellington, 34 pp. NNTT, 2013: National Native Title Tribunal Dataset: Determinations of Native Title, as per s. 192 Native Title Act 1993 (Cwlth), National Native Title Register - boundaries and core attributes (online dataset, accessed 31 August 2013, at www.ga.gov.au/meta/ANZCW0703011416.html). Geoscience Australia, Canberra. 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. Department of Climate Change and Energy Efficiency, Canberra, 60 pp. Norman, B., W. Steffen, R. Webb, A. Capon, W. Maher, C. Woodroffe, K. Rogers, R. Tanton, Y. Vidyattama, J. Lavis, H. Sinclair, B. Weir, 2012: South East Coastal Adaptation (SECA): Coastal urban climate futures in SE Australia from Wollongong to Lakes Entrance. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 130 pp. Norman-Lopez, A., S. Pascoe, 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. Natural Resources Commission, New South Wales Government, Sydney, 32 pp. Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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. Department of Sustainability, Environment, Water, Population and Communities, Canberra, 72 pp. Nursey-Bray, M., D. Fergie, V. Arbon, L. Rigney, R. Palmer, J. Tibby, N. Harvey, L. Hackworth, 2013: Community Based Adaptation to Climate Change: The Arabana. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 133 pp. NWC, 2009: Australian Water Reform 2009: Second Biennial Assessment of Progress in Implementation of the National Water Initiative. National Water Commission, Canberra, 288 pp. NWC, 2010: The Impacts of Water Trading in the Southern Murray-Darling Basin: An Economic, Social and Environmental Assessment. National Water Commission, Canberra, 237 pp. NWC, 2011: The National Water Initiative - Securing Australia's Water Future: 2011 Assessment. National Water Commission, Canberra, 354 pp. NZIER, 2003: Māori Economic Development: Te Ohanga Whakaketanga Māori. New Zealand Institute of Economic Research, Wellington, New Zealand, 116 pp. O'Connell, M., 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, 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. R07/4. Environment Canterbury, Christchurch, 75 pp. O'Leary, G., B. Christy, A. Weeks, C. Beverly, J. Nuttall, P. Riffkin, G. Fitzgerald, 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, 1518 November 2010, Lincoln, New Zealand [Dove, H., Culvenor, R.A. (eds.)], Australian Society of Agronomy, Gosford, NSW. O'Neill, S.J., J. Handmer, 2012: Responding to bushfire risk: the need for transformative adaptation. Environmental Research Letters, 7(1), 014018. Odeh, I.O.A., D.K.Y. Tan, 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), 165179. OECD, 2010: Sustainable Management of Water Resources in Agriculture. Organisation for Economic CoOperation and Development, Paris, 122 pp. OECD, 2011: Dairy. In: Agricultural Outlook 2011 [OECD, FAO (eds.)]. OECD Publishing, Paris, pp. 159-173. OECD, 2013a: OECD Factbook 2013: Economic, Environmental and Social Statistics. OECD Publishing, Paris, 300 pp. OECD, 2013b: Water Security for Better Lives. OECD Studies on Water, OECD Publishing, Paris, 173 pp. Ogden, J., L. Basher, 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., I. Webster, 2011: Water for the environment. In: Water: Science and Solutions for Australia [Prosser, I.P. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 119-134. Oxford Economics, 2009: Valuing the effects of Great Barrier Reef Bleaching. Great Barrier Reef Foundation, Newstead, Qld, 102 pp. Packman, D., D. Ponter, T. Tutua-Nathan, 2001: Climate Change Working Paper: Māori Issues. New Zealand Climate Change Programme, Wellington, New Zealand, 18 pp. Palutikof, J., M. Parry, M. Stafford-Smith, A.J. Ash, S.L. Boulter, M. Waschka, 2013: The past, present and future of adaptation: setting the context and naming the challenges. In: Climate Adaptation Futures [Palutikof, J., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 3-29. Park, S.E., S. Crimp, N.G. Inman-Bamber, Y.L. Everingham, 2010: Sugarcane. In: Adapting Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 85-99. Park, S.E., S. Crimp, A.M. Dowd, P. Ashworth, E. Mendham, 2011: Understanding climate change adaptation as a national challenge. Working Paper No. 8. CSIRO Climate Adaptation Flagship, Canberra, 32 pp. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Park, S.E., N.A. Marshall, E. Jakku, A.M. Dowd, S.M. Howden, E. Mendham, 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, B. Dong, 2007: Decadal to multidecadal variability and the climate change background. Journal of Geophysical Research-Atmospheres, 112(D18), D18115. Parsons Brinkerhoff, 2009: Energy network infrastructure and the climate change challenge. Energy Networks Association, Canberra, 129 pp. Peace, R., 2001: Social exclusion: a concept in need of definition? Social Policy Journal of New Zealand, 16, 17-36. Pearce, A., 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, 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, E. Yang, 2011: Improved estimates of the effect of climate change on NZ fire danger. Scion Client Report No. 18087 in conjunction with NIWA, Wellington. New Zealand Fire Service, Wellington, 84 pp. Pearce, M., E. Willis, 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., Nel, P., Righarts, M. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, NZ, pp. 9-36. Peel, M.C., T.A. McMahon, 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, I. van Putten, 2009: The east coast Tasmanian rock lobster fishery – vulnerability to climate change impacts and adaptation response options. Department of Climate Change, Canberra, 100 pp. Penny, G., D. King, 2010: Transfer of Renewable Energy Technology Systems to Rural Māori Communities. National Institute of Water and Atmospheric Research (NIWA), Auckland, 59 pp. Petheram, L., K.K. Zander, B.M. Campbell, C. High, N. Stacey, 2010: 'Strange changes': Indigenous perspectives of climate change and adaptation in NE Arnhem Land (Australia). Global Environmental Change, 20(4), 681692. Petheram, C., P. Rustomji, F. Chiew, W. Cai, T. McVicar, J. Vleeshouwer, T. Van Neil, L. Li, R. Donohue, R. Cresswell, J. Teng, 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. Petrone, K., J. Hughes, T. Van Niel, R. Silberstein, 2010: Streamflow decline in southwestern Australia, 19502008. Geophysical Research Letters, 37, L043102. Pettit, C.J., C.M. Raymond, B.A. Bryan, H. Lewis, 2011: Identifying strengths and weaknesses of landscape visualisation for effective communication of future alternatives. Landscape and Urban Planning, 100(3), 231241. Pham, T.D., D.G. Simmons, 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., W. Hill, 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., 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., J.G. Castley, 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, A.J. Hobday, 2010: Climate-driven range changes in Tasmanian intertidal fauna. Marine and Freshwater Research, 61(9), 963-970. Pittock, J., L.J. Hansen, R. Abell, 2008: Running dry: freshwater biodiversity, protected areas and climate change. Biodiversity, 9(3/4), 30-38. Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pittock, J., 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., 2013: Lessons from adaptation to sustain freshwater environments in the Murray–Darling Basin, Australia. Wiley Interdisciplinary Reviews: Climate Change, 4(5), 429-438. PMSEIC, 2010a: Australia and Food Security in a Changing World. Prime Minister’s Science, Engineering and Innovation Council, Canberra, 90 pp. PMSEIC, 2010b: Challenges at Energy-Water-Carbon Intersections. Prime Minister’s Science, Engineering and Innovation Council, Canberra, 88 pp. Polglase, P., A. Reeson, C. Hawkins, K. Paul, A. Siggins, J. Turner, D. Crawford, T. Jovanovic, T. Hobbs, K. Opie, J. Carwardine, A. Almeida, 2011: Opportunities for carbon forestry in Australia: Economic assessment and constraints to implementation. CSIRO, Canberra, 30 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, 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, 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, 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. IAHS Red Book 344 [Franks, S.W., Boegh, E., Blyth, E., Hannah, D.M., Yilmaz, K.K. (eds.)]. International Association of Hydrological Sciences, 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, 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., F. Chiew, 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. Potter, N.J., 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. Power, S., F. Tseitkin, S. Torok, B. Lavery, R. Dahni, 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, 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, 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, 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., R.N. Jones, 2008: Screening Climatic and Non-Climatic Risks to Australian Catchments. Geographical Research, 46(3), 258-274. Preston, B., T.F. Smith, C. Brooke, R. Gorddard, T. Measham, G. Withycombe, B. Beveridge, C. Morrison, K. McInnes, D. Abbs, 2008: Mapping climate change vulnerability in the Sydney Coastal Councils Group. Sydney Coastal Councils Group and Australian Government, Sydney, 124 pp. Preston, B., M. Stafford-Smith, 2009: Framing vulnerability and adaptive capacity assessment: Discussion paper. CSIRO Climate Adaptation Flagship Working paper No. 2. CSIRO, Canberra. Preston, B., C. Brooke, T. Measham, T. Smith, 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, 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. Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pride, S., B. Frame, D. Gill, 2010: Futures Literacy in New Zealand. Journal of Futures Studies, 15(1), 135-144. Prober, S.M., 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. Final Inquiry Report No. 46. Productivity Commission, Melbourne, 431 pp. Productivity Commission, 2012: Barriers to effective climate change adaptation. Inquiry Report No. 59. Productivity Commission, Canberra, 385 pp. Prosser, I.P., I.D. Rutherfurd, J.M. Olley, W.J. Young, P.J. Wallbrink, 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., 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, 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, Sydney, NSW, 6 pp. QFCI, 2012: Queensland Floods Commission of Inquiry Final Report. Queensland Floods Commission of Inquiry, Brisbane, Qld, 623 pp. QRC, 2011: Queensland Resources Council, submission to Queensland floods commision of inquiry. Queensland Resources Council, Brisbane, Qld, 43 pp. Queensland Government, 2011: Understanding Floods: Questions and Answers. Queensland Government, Brisbane, 36 pp. Quiggin, J., D. Adamson, P. Schrobback, S. Chambers, 2008: The Implications for Irrigation in the Murray-Darling Basin. Paper prepared for the Garnaut Climate Change Review. Garnaut Climate Change Review, Canberra, 59 pp. Quiggin, J., D. Adamson, S. Chambers, 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, 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, A. Elmahdi, 2013b: A biophysical and economic model of agriculture and water in the Murray-Darling Basin, Australia. Environmental Modelling & Software, 41(0), 98-106. QUT, 2010: Impacts and Adaptation Responses of Infrastructure and Communities to Extreme Events. The Southern Australian Experience of 2009. Prepared by Queensland University of Technology. National Climate Change Adaptation Research Facility, Gold Coast, Qld. Radcliffe, J.E., J.A. Baars, 1987: Temperate pasture production. In: Ecosystems of the World [Snaydon, R.W. (eds.)]. Elsevier, Amsterdam, pp. 7-17. Rafter, A., 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 (Centre for Australian Weather and Climate Research), Vol 3 [P. A. Sandery, T. Leeuwenburg, G. Wang, A. J. Hollis (eds.)], 44-49. Rajanayaka, C., J. Donaggio, 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. Ministry for the Environment, Wellington, 118 pp. Randall, A., T. Capon, T. Sanderson, D. Merrett, G. Hertzler, 2012: Making decisions under the risks and uncertainties of future climates. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 72 pp. Rau, G.H., E.L. McLeod, O. Hoegh-Guldberg, 2012: The need for new ocean conservation strategies in a highcarbon dioxide world. Nature Climate Change, 2(10), 720-724. Raymond, C.M., 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. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Raymond, C.M., G. Brown, 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. Raymond, C.M., J. Spoehr, 2013: The acceptability of climate change in agricultural communities: Comparing responses across variability and change. Journal of Environmental Management, 115, 69-77. RBA, 2006: Statement on Monetary Policy – November 2006. Reserve Bank of Australia, Canberra, 59 pp. RBA, 2011: Statement on Monetary Policy. Box B: An update on the impact of natural disasters in Queensland. Reserve Bank of Australia, Canberra, 5 pp. Reijnders, L., 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, 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., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 26-43. Reisinger, A., D. Wratt, S. Allan, 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., Berrang-Ford, L. (eds.)]. Springer, Toronto, Canada, pp. 303-319. Reisinger, A., P. Havlik, K. Riahi, O. Vliet, M. Obersteiner, 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, R. Chapman, 2014: 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., Kaye, R., Kelly, M., Travers, A. (eds.)]. CRC Press, London, www.crcpress.com/product/isbn/9780415464871. Ren, Z., X. Wang, Z. Chen, 2012: Energy efficient houses for heat stress mitigation: 2009 Melbourne heatwave scenarios. In: 10th International Healthy Building Conference, 8-12 July 2012 [International Society of Indoor Air Quality and Climate (eds.)], Brisbane, Qld. pp 4D.5. Renwick, J., P. Mladenov, J. Purdie, A. McKerchar, 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., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 70-81. Reser, J.P., 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, M. Ellul, 2011: The threat of climate change: psychological response, adaptation, and impacts. In: Climate change and human well being. International and Cultural Psychology Series [Weissbecker, I. (eds.)]. Springer, New York, pp. 19-42. Reser, J.P., G.L. Bradley, 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., Grimaldo, V. (eds.)]. Nova Science Publishers, New York, pp. 1-34. Reser, J.P., G.L. Bradley, A.I. Glendon, M.C. Ellul, R. Callaghan, 2012b: Public risk perceptions, understandings and responses to climate change and natural disasters in Australia, 2010 and 2011. Final report. Griffith University, School of Psychology, Gold Coast, Qld, 246 pp. Reser, J.P., G.L. Bradley, A.I. Glendon, M.C. Ellul, R. Callaghan, 2012c: Public Risk Perceptions, Understandings, and Responses to Climate Change and Natural Disasters in Australia and Great Britain. Griffith University, School of Psychology, Gold Coast, Qld, 298 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, 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. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 230 pp. Rhodes, B.G., J.E. Fagan, K.S. Tan, 2012: Responding to a rapid climate shift - experiences from Melbourne, Australia. In: World Congress on Water, Climate and Energy, 13-18 May 2012 [International Water Association (eds.)], Dublin, Ireland. pp 8. Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Richardson, A.J., E.S. Poloczanska, 2009: Australia's Oceans. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 6 pp. Richardson, A.J., D. McKinnon, K.M. Swadling, 2009: Zooplankton. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 15 pp. Rickards, L., 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. Risbey, J.S., M.J. Pook, P.C. McIntosh, M.C. Wheeler, H.H. Hendon, 2009: On the Remote Drivers of Rainfall Variability in Australia. Monthly Weather Review, 137(10), 3233-3253. Ritchie, E.G., 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., T. Weeks, 2011: Adaptation to climate change in New Zealand. In: Climate Change Law and Policy in New Zealand [Cameron, A. (eds.)]. LexisNexis, Wellington, pp. 345-392. Roberts, D.A., E.L. Johnston, 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. 15th Agronomy Conference, 15-18 November 2010, Lincoln, New Zealand [Dove, H., Culvenor, R.A. (eds.)], Australian Society of Agronomy, Gosford, NSW. Rogan, R., M. O’Connor, 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, 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, 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., Nel, P., Righarts, M. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, NZ, pp. 93-100. Roman, C.E., A.H. Lynch, 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, 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, V. Cárdenas, 2011: The impact of weather variability on British outbound flows. Climatic Change, 105(1), 281-292. 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, 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, 108(1), 220-225. Rouse, H.L., 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., 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, 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, 402 pp. Russell, B.D., J.A.I. Thompson, L.J. Falkenberg, S.D. Connell, 2009: Synergistic effects of climate change and local stressors: CO2 and nutrient-driven change in subtidal rocky habitats. Global Change Biology, 15(9), 21532162. Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Russell, S.L., A. Greenaway, F. Carswell, S. Weaver, 2013: Moving beyond “mitigation and adaptation”: examining climate change responses in New Zealand. Local Environment, advance online, 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, 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, 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. Landcare Research, Wellington, New Zealand, 130 pp. Sadras, V.O., 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, A.B. Mullan, 2001: Interdecadal Pacific Oscillation and South Pacific climate. International Journal of Climatology, 21(14), 1705-1721. Salinger, M.J., G. Griffiths, 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, M. Houghton, 2008: Impact of bushfires on tourism and visitation in alpine national parks. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 42 pp. Sato, Y., D.G. Bourne, 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 BBiological Sciences, 276(1668), 2795-2803. Sattler, P., M. Taylor, 2008: Building Nature's Safety Net 2008: Progress on the directions for the National Reserve System. WWF-Australia Report. WWF-Australia, Sydney, 64 pp. Saunders, C., W. Kaye-Blake, L. Marshall, S. Greenhalgh, 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, J. Turner, 2010: Modelling Climate Change Impacts on Agriculture and Forestry with the Extended LTEM (Lincoln Trade and Environment Model). Ministry of Agriculture and Forestry, Wellington, NZ, 31 pp. Saunders, D.A., P. Mawson, 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 report on the inquiry into climate change and environmental impacts on coastal communities. Parliament of Australia, Canberra, 368 pp. Scealy, B., D. Newth, D. Gunasekera, 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), 3351. Schiff, A., S. Becken, 2011: Demand Elasticities for Tourism in New Zealand. Tourism Management, 32(3), 564575. Schofield, N., 2011: Climate Change and its Impacts - current understanding, future directions. In: Basin futures: water reform in the Murray-Darling Basin [Grafton, Q., Connell, D. (eds.)]. ANU E-Press, Canberra, pp. 81100. Schrobback, P., D. Adamson, 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, 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., Boulter, S., Ash, A., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 133-140. Scott, D., S. Gössling, C.M. Hall, 2012: International tourism and climate change. Wiley Interdisciplinary Reviews: Climate Change, 3(3), 213-232. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 SCRGSP, 2011: Overcoming Indigenous disadvantage: Key indicators 2011. Steering Committee for the Review of Government Service Provision, Productivity Commission, Canberra, 64 pp. Scutella, R., R. Wilkins, W. Kostenko, 2009: Estimates of Poverty and Social Exclusion in Australia: A Multidimensional Approach. Melbourne Institute of Applied Economic and Social Research Working Paper No. 26/09. Melbourne University, Melbourne, 67 pp. Serrao-Neumann, S., F. Crick, B. Harman, M. Sano, O. Sahin, R. Staden, G. Schuch, S. Baum, D. Low Choy, 2013: Improving cross-sectoral climate change adaptation for coastal settlements: insights from South East Queensland, Australia. Regional Environmental Change, advance online, DOI: 10.1007/s10113-013-0442-6. SGS, 2012: Tasmanian Coastal Adaptation Pathways Project. Lauderdale Recommended Actions. Report prepared by SGS Economics & Planning. Local Government Association of Tasmania, Hobart, 55 pp. Shakesby, R.A., P.J. Wallbrink, S.H. Doerr, P.M. English, C.J. Chafer, G.S. Humphreys, W.H. Blake, 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. New Zealand Business Council for Sustainable Development, Auckland, 34 pp. Sharma, V., S.v.d. Graaff, B. Loechel, D. Franks, 2013: Extractive resource development in a changing climate. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 116 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: State of Australian Cities National Conference, 29 November-2 December 2011, Melbourne, Vic [Whitzman, C., Fincher, R. (eds.)], Faculty of the Built Environment, University of New South Wales, Sydney, 15 pp. Sheppard, C.S., 2012: Potential spread of recently naturalised plants in New Zealand under climate change. Climatic Change, advance online, doi: 10.1007/s10584-012-0605-3. Shoo, L.P., C. Storlie, J. Vanderwal, J. Little, 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, 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. 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, NZ, 68 pp. Singh, S., S. Davey, 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 International Adaptation Forum on Climate Change Impacts on Water Supply, Washington DC, January, 2010. Melbourne Water, Melbourne, 24 pp. Smale, D.A., T. Wernberg, 2013: Extreme climatic event drives range contraction of a habitat-forming species. Proceedings of the Royal Society of London, Series B: Biological Sciences, 280(1754), 20122829. Smales, I., 2006: Wind farm collision risk for birds. Cumulative risks for threatened and migratory species. Report by Biosis Research Pty Ltd. Department of Environment and Heritage, Canberra, 237 pp. Smart, A., A. Aspinall, 2009: Water and the electricity generation industry. Waterlines report No. 18. National Water Commission, Canberra, 115 pp. Smart, R.E., 2010: A lump of coal, a bunch of grapes. Journal of Wine Research, 21(2), 107-111. Smart, G.M., A.I. McKerchar, 2010: More flood disasters in New Zealand. Journal of Hydrology (New Zealand), 49(2), 69-78. Smith, T.F., C. Brooke, T.G. Measham, B. Preston, R. Gorddard, G. Withycombe, B. Beveridge, C. Morrison, 2008: Case studies of adaptive capacity: systems approach to regional climate change adaptation strategies. Sydney Coastal Councils Group Inc, Sydney, NSW. Smith, I.N., B. Timbal, 2010: Links between tropical indices and southern Australian rainfall. International Journal of Climatology, 32(1), 33-40. 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, C. Stephenson, 2010: Towards Enhancing Adaptive Capacity for Climate Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Change Response in South East Queensland. The Australasian Journal of Disaster and Trauma Studies, 2010-1, online. Smith, H.G., G.J. Sheridan, P.N.J. Lane, P. Nyman, S. Haydon, 2011a: Wildfire effects on water quality in forest catchments: A review with implications for water supply. Journal of Hydrology, 396(1–2), 170-192. Smith, W., C. Davies-Colley, A. Mackay, G. Bankoff, 2011b: Social impact of the 2004 Manawatu floods and the ‘hollowing out’ of rural New Zealand. Disasters, 35(3), 540−553. 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, R.W. Carter, 2014: 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., Kay, R.C., Kelly, M., Travers, A. (eds.)]. CRC Press, London, www.crcpress.com/product/isbn/9780415464871. SNZ, 2010a: National ethnic population projections: 2006 (base) - 2026 update. Statistics New Zealand, Wellington, 31 pp. SNZ, 2010b: Are New Zealanders living closer to the coast? Internal Migration. Statistics New Zealand, Wellington, 4 pp. SNZ, 2010c: Subnational Population Projections: 2006 (base) – 2031 update. Statistics New Zealand, Wellington, 15 pp. SNZ, 2011: National Accounts: Long-term data series. Table E.1.2 real gross domestic product (Excel data spreadsheet, online at www.stats.govt.nz). Statistics New Zealand, Wellington. SNZ, 2012a: National Population Projections: 2011 (base) – 2061. Statistics New Zealand, Wellington, 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, Wellington, NZ, 199 pp. SNZ, 2013: New Zealand's seafood industry - Infographic. Statistics New Zealand, Wellington, 1 pp. 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 Committee, Canberra, 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. Spickett, J., H. Brown, D. Katscherian, 2008: Health impacts of climate change: Adaptation strategies for Western Australia. Western Australia Department of Health, Perth, 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, 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., L. Horrocks, A. Harvey, C. Hamilton, 2011: Rethinking adaptation for a 4°C world. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 196-216. Stafford-Smith, M., 2013: Scenarios for picturing a future adapted to +4°C. In: Climate Adaptation Futures [Palutikof, J., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 119-125. Stamatov, V., 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, 47 pp. Stark, C., K. Penney, A. Feng, 2012: 2012 Australian Energy Update. Bureau of Resources and Energy Economics, Canberra, 26 pp. Steffen, W., A.A. Burbidge, L. Hughes, R. Kitching, D. Lindenmayer, W. Musgrave, M. Stafford-Smith, P.A. Werner, 2009: Australia's biodiversity and climate change: A strategic assessment of the vulnerability of Australia's biodiversity to climate change. A report to the Natural Resource Management Ministerial Council commissioned by the Australian Government. CSIRO Publishing, Collingwood, Vic, 289 pp. Steffen, W., J. Sims, J. Walcott, G. Laughlin, 2011: Australian agriculture: Coping with dangerous climate change. Regional Environmental Change, 11(Suppl. 1), 205-214. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Stehlik, D., G. Lawrence, 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, P. Thorsnes, 2010: Energy cultures: A framework for understanding energy behaviours. Energy Policy, 38(10), 6120-6129. Stewart, H.T.L., D.H. Race, A.L. Curtis, 2011: New Forests in Changing Landscapes in South-East Australia. International Forestry Review, 13(1), 67-79. Stewart, M.G., X. Wang, 2011: Risk Assessment of Climate Adaptation Strategies for Extreme Wind Events in Queensland. CSIRO, Canberra, 85 pp. Stewart, M.G., X. Wang, M.N. Nguyen, 2012: Climate change adaptation for corrosion control of concrete infrastructure. Structural Safety, 35(0), 29-39. Stokes, C., A. Ash, 2007: Impacts of climate change on marginal tropical animal production systems. In: Agroecosystems in a Changing Climate [Newton, P.C.D., Carran, R.A., Edwards, G.R., Niklaus, P.A. (eds.)]. CRC Press, New York, pp. 323-328. Stork, N.E., J. Balston, G.D. Farquhar, P.J. Franks, J.A.M. Holtum, 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., C. Maller, 2011: Integrating health, housing and energy policies: Social practices of cooling. Building Research and Information, 39(Compendex), 154-168. Stroombergen, A., A. Tait, J. Renwick, K. Patterson, 2006: The relationship between New Zealand's climate, energy, and the economy in 2025. Kotuitui – New Zealand Journal of Social Sciences Online, 1, 139-160. Stroombergen, A., 2010: The International Effects of Climate Change on Agricultural Commodity Prices, and the Wider Effects on New Zealand. Working Paper 10-14. Motu Economic and Public Policy Research, Wellington, New Zealand, 37 pp. Suncorp, 2013: Risky Business. Insurance and Natural Disaster Risk Management. Suncorp Personal Insurance Public Policy. Suncorp, Brisbane, Qld, 20 pp. Sutherst, R.W., B.S. Collyer, 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), 467480. Sutton, P.J.H., M. Bowen, 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 2050. Bureau of Resources and Energy Economics, Canberra, 66 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, 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, J. Reid, 2008b: Use of climate, soil, and tree species information for identifying land use options across the Gisborne district. National Institute of Water and Atmospheric Research (NIWA), Wellington, 104 pp. Taptiklis, N., 2011: Climate resilient water management in Wellington, New Zealand. NZCCRI 2011 Report 09. New Zealand Climate Change Research Institute, Victoria University of Wellington, Wellington, 62 pp. Taylor, M.A.P., 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, H. van Roon, 2011: Building capacity in urban sustainability assessment through use of a scenarios game. Journal of Education for Sustainable Development, 5(1), 75-87. Taylor, B.M., B.P. Harman, S. Heyenga, R.R.J. McAllister, 2012a: Property Developers and Urban Adaptation: Conceptual and Empirical Perspectives on Governance. Urban Policy and Research, 30(1), 5-24. Taylor, S., L. Kumar, N. Reid, 2012b: 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, P. Johnstone, 2012: Broadacre cropping: Adapting broad acre cropping to climate change. In: Enhanced climate Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 change impact and adaptation evaluation: A comprehensive analysis of New Zealand's land-based primary sectors. [Clark, A., Nottage, R. (eds.)]. Ministry for Primary Industries, New Zealand, pp. 189-236. Teng, J., F. Chiew, J. Vaze, S. Marvanek, 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, advance online, doi:10.1175/JHM-D-11-097.1. 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, Geneva, Switzerland, 152 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. Thomas, D.T., J. Sanderman, S.J. Eady, D.G. Masters, P. Sanford, 2012: Whole Farm Net Greenhouse Gas Abatement from Establishing Kikuyu-Based Perennial Pastures in South-Western Australia. Animals, 2(3), 316330. Thompson, D., J. Wallace, 2000: Annular modes in the extratropical circulation. Part I: Month-to-month variability. Journal of Climate, 13(5), 1000-1016. Thompson, P.A., M.E. Baird, T. Ingleton, 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, 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, M. Battaglia, 2012: Australia and New Zealand Perspectives on Climate Change and Agriculture. In: Agriculture's Contributions to Climate Change Solutions: Mitigation and Adaptation at Global and Regional Scales. Handbook of Climate Change and Agroecosystems, Vol 2 [Rosenzweig, C., Hillel, D. (eds.)]. American Society of Agronomy and Imperial College Press, New York, pp. 107-141. Thresher, R.E., J.A. Koslow, A.K. Morison, 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., J.M. Arblaster, S. Power, 2006: Attribution of the late-twentieth-century rainfall decline in southwest Australia. Journal of Climate, 19(10), 2046-2062. Timbal, B., 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. Arblaster, K. Braganza, E. Fernandez, H. Hendon, B. Murphy, M. Raupach, C. Rakich, I. Smith, K. Whan, 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, Melbourne, Vic, 202 pp. Timbal, B., R. Kounkou, 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, 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., Jack, M. (eds.)]. Scion, Rotorua, NZ, 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, 18 pp. Tong, S.L., C.Z. Ren, 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, A.G. Barnett, 2010b: Assessment of Heat-Related Health Impacts in Brisbane, Australia: Comparison of Different Heatwave Definitions. Plos One, 5(8), e12155. Tourism Queensland, 2007: Regional Tourism Crisis Management Plan Template. A guide to preparing a Regional Tourism Crisis Management Plan. Tourism Queensland, Brisbane, Qld, 30 pp. Tourism Queensland, 2010: Climate Futures Industry Tool. Tourism Queensland, Brisbane, Qld, 31 pp. Tourism Victoria, 2010: Crisis Essentials. Crisis Management for tourism businesses. Tourism Victoria, Melbourne, Vic, 18 pp. TPK, 2007: A time for change in Māori economic development. Te Puni Kokiri, Wellington, New Zealand, 36 pp. TRA, 2010: Impact of the drought on tourism in the Murray River region: Summary of results. Tourism Research Australia, Department of Resources, Energy and Tourism, Canberra, 3 pp. Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Traill, L.W., K. Perhans, C.E. Lovelock, A. Prohaska, S. McFallan, J.R. Rhodes, 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., 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, 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 Council's story. Waitakere City Council, Waitakere, 56 pp. Troy, P. (eds.), 2008: Troubled waters: confronting the water crisis in Australia’s cities. ANU E-Press, Canberra, 217 pp. Tryhorn, L., 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, Darwin, 20 pp. Turton, S., W. Hadwen, R. Wilson, B. Jorgensen, 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, 77 pp. Turton, S., T. Dickson, W. Hadwen, B. Jorgensen, T. Pham, D. Simmons, P. Tremblay, 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, 2012: World Urbanization Prospects. The 2011 Revision. Department of Economic and Social Affairs, United Nations Secretariat, New York, 318 pp. Valenzuela, E., 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., 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, 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, L. Zhang, 2006: Risks to shared water resources of the Murray-Darling Basin. Publication 22/06. Murray Darling Basin Commission, Canberra, 49 pp. Vaneckova, P., P.J. Beggs, R.J. de Dear, 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, Melbourne, 42 pp. Veland, S., R. Howitt, 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, S.W. Franks, 2004: Multidecadal variability of rainfall and streamflow: Eastern Australia. Water Resources Research, 40(10), W10201. Verdon-Kidd, D.C., A.S. Kiem, 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), Gold Coast, Qld, 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, A.D. Rogers, 2009: The coral reef crisis: The critical importance of <350 ppm CO2. Marine Pollution Bulletin, 58(10), 1428-1436. Verschuuren, J., 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, 16 pp. Victorian Government, 2009b: Heatwave Plan for Victoria 2009-2010: Protecting health and reducing harm from heatwaves. Department of Health, Melbourne, Vic, 42 pp. Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Walker, S.J., T.A. Schlacher, 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, J.F. Weltzin, 2007: Responses of soil respiration to elevated CO2, air warming, and changing soil water availability in an old-field grassland. Global Change Biology, 13(11), 2411-2424. Wang, X., D. Chen, 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, 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. Climate Adaptation Flagship Working Paper No. 6. CSIRO, Brisbane, 28 pp. Wang, X., R. McAllister, 2011: Adapting to Heatwaves and Coastal Flooding. In: Climate Change: Science and Solutions for Australia [Cleugh, H., Stafford-Smith, M., Battaglia, M., Graham, P. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 73-84. Wang, X., M. Stewart, M. Nguyen, 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, P.C.D. Newton, 2013: Reduced nifH gene and transcript numbers indicate a reduced capacity for nitrogen fixation in nodules of white clover after exposure of a grazed grassland to elevated CO2. Biogeosciences Discussion, 10, 9867-9896. 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, C.D. Buxton, 2012: Ecosystem model of Tasmanian waters explores impacts of climate change induced changes in primary productivity. Ecological Modelling, advance online, doi: 10.1016/j.ecolmodel.2012.05.008. 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, 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. Scion, Rotorua, 149 pp. Watt, M.S., D.J. Palmer, 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, 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, Wellington, 39 pp. Webb, L.B., P.H. Whetton, 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, E.W.R. Barlow, 2012a: 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, J.C. Scanlan, 2012b: Interacting effects of vegetation, soils and management on the sensitivity of Australian savanna rangelands to climate change. Climatic Change, 112(3-4), 925-943. Webb, B., 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), Gold Coast, Qld, 84 pp. Webb, B., R. McKellar, R. Kay, 2013: Climate change adaptation in Australia: experience, challenges and capacitybuilding. Australian Journal of Emergency Management, advance online, DOI: 10.1080/14486563.2013.835285. Weber, E.P., A. Memon, 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, 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 [CSIRO Sustainable Agriculture Flagship (eds.)]. Department of Infrastructure, Transport, Regional Development and Local Government, Canberra, pp. 10-1 to 10-53. Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Welbergen, J.A., S.M. Klose, N. Markus, P. Eby, 2008: Climate change and the effects of temperature extremes on Australian flying-foxes. Proceedings of the Royal Society B-Biological Sciences, 275(1633), 419-425. Wernberg, T., B.D. Russell, M.S. Thomsen, C.F.D. Gurgel, C.J.A. Bradshaw, E.S. Poloczanska, 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, 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., 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., 2012: Adapting to Climate Change – Revising our Approach to Estimating Future Floods. University of New South Wales, Sydney, 25 pp. Westra, S., J. Evans, R. Mehrotra, 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, 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, D.A., D.S. Crombie, J. Kinal, M. Battaglia, J.F. McGrath, D.S. Mendham, 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., 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., Wilkes, K. (eds.)]. Earthscan Publications, Oxford, pp. 117-133. Whitehead, P.J., P. Purdon, P.M. Cooke, J. Russell-Smith, 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., Whitehead, P.J., Cooke, P.M. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 287-312. Whittaker, J., J. Handmer, D. Karoly, 2013: After 'Black Saturday': adapting to bushfires in a changing climate. In: Natural Disasters and Adaptation to Climate Change [Boulter, S., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 115-136. Wilby, R.L., 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, 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, R.J., R.A. Bradstock, G.J. Cary, N.J. Enright, A. Gilll, M,, A.C. Leidloff, C. Lucas, R. Whelan, A.N. Andersen, D.M.J.S. Bowman, P.J. Clarke, G.D. Cook, K. Hennessy, A. York, 2009: Interactions between climate change, fire regimes and biodiversity in Australia - a preliminary assessment. Department of Climate Change, Canberra, 196 pp. Williams, S.E., L.P. Shoo, J.L. Isaac, A.A. Hoffmann, G. Langham, 2008: Towards an Integrated Framework for Assessing the Vulnerability of Species to Climate Change. Plos Biology, 6(12), 2621-2626. Williams, N.S.G., J.P. Rayner, 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. 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, A. Tait, 2007: Climate change and the New Zealand marine environment. NIWA Client Report NEL2007-025. Department of Conservation, Wellington, 81 pp. Willsman, A., T. Chinn, J. Hendrikx, A. Lorrey, 2010: New Zealand Glacier Monitoring: End of Summer Snowline Survey 2010. NIWA Client Report no. CHC2010-113. National Institute of Water and Atmospheric Research (NIWA), Christchurch, 132 pp. 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, J., 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, R., 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), Gold Coast, 124 pp. Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wilson, R., 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., Phillips, M. (eds.)]. CAB International, Wallingford, pp. 233-253. Winterbourn, M.J., S. Cadbury, C. Ilg, 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., 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, J. Diettrich, 2008: Projected climate and river flow for the Rangitata catchment for 2040. NIWA Client Report: CHC2008-097. National Institute of Water and Atmospheric Research (NIWA), Christchurch, 28 pp. Woodward, A., S. Hales, N. de Wet, 2001: Climate change: potential effects on human health in New Zealand. A report prepared for the Ministry for the Environment as part of the New Zealand Climate Change Programme. Ministry for the Environment, Wellington, New Zealand, 27 pp. Woodward, E., S. Jackson, M. Finn, 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, 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, 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. Ministry of Agriculture and Forestry, Wellington, 112 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, 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, 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., 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, 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, advance online, doi:10.1007/s11027-012-9376-4. Zammit, C., 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, 46 pp. Zander, K.K., L. Petheram, 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., N. Beaumont, 2011: Green Tourism Futures: Climate Change Responses by Australian Government Tourism Agencies. In: CAUTHE 2011: National Conference: Tourism: Creating a Brilliant Blend, 8-11 February 2011, Adelaide, SA. pp 850-865. Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 Chapters 3, 6 (ocean changes, including acidification), 11, 12 (projections), 13 (sea level), and 14 (regional climate phenomena). (*) medium confidence, (**) high confidence, (***) very high confidence, (****) virtually certain Climate variable Mean air temperature Observed change Direction of projected change Aus: Increased by 0.09 ± 0.03°C per decade since 19111 (***) Aus and NZ: Increase3-8 (****); greatest over inland Aus and least in coastal areas and NZ5-8 (***) NZ: Increased by 0.09 ± 0.03°C per decade since 19092 (***) Examples of projected magnitude of change (relative to ~1990, unless otherwise stated) Aus: 0.6-1.5°C (2030 A1B), 1.0-2.5°C (2070 B1), 2.2-5.0°C (2070 A1FI)3 NZ: 0.3-1.4°C (2040 A1B), 0.7-2.3°C (2090 B1), 1.6-5.1°C (2090 A1FI)5 CMIP5 RCP4.5, rel. to ~19959: N Aus: 0.3-1.6°C (2016-2035), 0.7-2.6°C (2046-2065) Sea surface temperature Aus: Increased by about 0.12°C per decade for NW&NE Aus and by about 0.2°C per decade for SE Aus since 195014,15 (***) Aus and NZ: Increase3,7,8 (***) with greater increase in the Tasman sea region (*) 3,7 NZ: Increased by about 0.07°C per decade over 1909-20092 (***) Air temperature extremes Precipitation Aus and NZ: Significant trend since 1950: cool extremes have become rarer and hot extremes more frequent and intense16-19 (**).The Australian heatwave of 2012/13 was exceptional in heat, duration and spatial extent 20 . Aus and NZ: Hot days and nights more frequent and cold days and cold nights less frequent during the 21st century3,5,21-24 (**) Aus: Late autumn/winter decreases in SW Aus since the 1970s and in SE Aus since the mid 1990s, and annual increases in NW Aus since the 1950s29-31 (***) Aus: Annual decline in SW Aus (**), elsewhere on most of the southern (*) and NE (low confidence) continental edges, with reductions strongest in the winter half year3,4,9,33-35 (**). Direction of annual change elsewhere is uncertain3,35,36 (Figure 25.1) (**) NZ: Spring and autumn frost-free land to at least triple by 2080s24; up to 60 more hot days (>25°C max.) for northern areas by 20905 Aus: For 2030 A1B, annual changes of-10% to +5% (N Aus) and -10% to 0% (S Aus), for 2070 B1, -15% to +7.5% (N&E Aus) and 15% to 0% (S Aus), and for 2070 A1FI, -30% to +20% (N&E Aus) and -30% to +5% (S Aus), with larger changes seasonally3 Precipitation extremes Aus: Indices of annual daily extremes (e.g. 95th and 99th percentile rainfalls) show mixed or insignificant trends21,48, but significant increase is evident in recent decades for shorter duration (sub-daily) events49,50 (**) NZ: Extreme annual 1-day rainfall decrease in north and east and increase in west since 193032 (*) Subject to Final Copyedit NZ: In the South Island, annual increase in the west and south and decrease in north-east. In the North Island, increase in the west and decrease in eastern and northern regions 5,34,37 (Figure 25.1) (*) Aus and NZ: 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 99 percentile daily extremes (low confidence)5,8,21,51-56 85 Aus: A significant contribution to observed change attributed to anthropogenic climate change10 (**) with some regional variations attributed to atmospheric circulation variations11,12 NZ: Observed change partially attributed to anthropogenic climate change13 (*) S Aus & NZ: 0.1-1.0°C (2016-2035), 0.61.7°C (2046-2065) Aus: 0.6-1.0°C (2070 B1) and 1.6-2.0°C (2070 A1FI) for southern coastal and 1.21.5°C (2070 B1) and 2.2-2.5°C (2070 A1FI) elsewhere3 NZ: Similar to projected changes in mean air temperature for coastal waters5 Aus: Hot days in Melbourne (>35°C max.) increase by 20-40% (2030 A1B), 30-90% (2070 B1) and 70-190% (2070 A1FI)3 NZ: 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 north-east of the South Island and east and north of the North Island32 (***) Additional comments NZ: For 2040 A1B, annual changes of -5% to +15% (S&W) and -15% to +10% (N&E) and for 2090 A1B, -10% to +25% (S&W) and 20% to +15% (N&E) based on downscaled projections with larger changes seasonally5,37 Aus: For 2090 A2, CMIP3 give increases in the intensity of the 20 year daily extreme of around +200% to -25% depending on region and model52 NZ: Increases of daily extreme rainfalls of around 8% per degree C are projected but with significant regional variations5,56 Aus: Observed trends partly attributable to anthropogenic climate change (**) as they are consistent with mean warming and historical simulations18,19,21,25, although other factors may have contributed to high extremes during droughts26-28 Aus: Observed decline in SW is related to atmospheric circulation changes38-40 (***), other factors41, and partly attributable to anthropogenic climate change40-43 (**). The recent SE rainfall decline is also related to circulation changes31,44-46 (**), with some evidence of an anthropogenic component47 NZ: Observed trends related to increased westerly winds32. Projected annual trends dominated by winter and spring trends related to increased westerlies5 Aus and NZ: The sign of observed trends mostly reflects trends in mean rainfall (e.g. there is a decrease in mean and daily extremes in SW Aus)21,32,49. Similarly, future increases in intensity of extreme daily rainfall are more likely where mean rainfall is projected to increase3,5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Drought Aus: Defined using rainfall only, drought occurrence over the period 1900-2007 has not changed significantly57 (**) Aus and NZ: Drought frequency is projected to increase in southern Australia 8,54,57,59,60 (*) and in many regions of New Zealand58,61 (*) NZ: Defined using a soil water balance model, there has been no trend in drought occurrence since 197258 (*) Winds Aus: Significant decline in storminess over SE Aus since 188564 (*), but inconsistent trends in wind observations since 197565,66 NZ: Mean westerly flow increased during the late 20th century (1978–1998), associated with the positive phase of the IPO67,68 Mean sea level Aus: From 1900-2011 the average rate of relative sea level rise (SLR) was 1.4±0.6 mm/yr72 (***) NZ: The average rate of relative SLR was 1.7±0.1 mm/yr over 1900-200973 (***) Aus: Increases in winds in 20-30°S band, with little change to decrease elsewhere, except for winter increases over Tasmania. Decrease to little change in extremes (99th percentile) over most of Australia except Tasmania in winter69 (*) NZ: Mean westerly winds and extreme winds (based on projected changes in circulation patterns) are projected to increase, especially in winter5,70 (*) Aus and NZ: Regional sea level rise will very likely exceed the 1971-2000 historical rate, consistent with global mean trends74. Mean sea level will continue to rise for at least several more centuries74 (***) Extreme sea level Aus and NZ: Extreme sea levels have risen at a similar rate to global SLR79 Aus and NZ: Projected mean SLR will lead to large increases in the frequency of extreme sea level events (***), with other changes in storm surges playing a lesser role80-83 Fire weather Aus: Increased since 1973(**) with 24 out of 38 sites showing increases in the 90th percentile of the McArthur Forest Fire Danger index84 Aus: Fire weather is expected to increase in most of southern Australia due to hotter and drier conditions (**), based on explicit model studies carried out for SE Australia 85-88, and change little or decrease in NE88 (*) NZ: Fire danger index is projected to increase in many areas89 (*) Tropical cyclones and other severe storms Aus: No regional change in the number of tropical cyclones (TCs) or in the proportion of intense TCs over 1981-200790 (*), but frequency of severe landfalling TCs in NE Aus has declined significantly since the late 19th Century91 and east-west distribution changed since 1980.92 There has been no trend in environments suitable for severe thunderstorms93 Subject to Final Copyedit Aus: Tropical cyclones are projected to increase in intensity and stay similar or decrease in numbers9,94, and occur further south94 (low confidence) NZ: Projected increase in the average intensity of cyclones in the south during winter, but a decrease elsewhere70 (*) 86 Aus: Occurrence under 2070 A1B and A2 ranges from a halving to 3 times more frequent in N. Aus, and 0-5 times more frequent in southern Aus60 NZ: Time spent in drought in eastern and northern New Zealand is projected to double or triple by 204061 Aus: Magnitude of simulated mean changes may exceed 10% under A1B for 2081-2100 relative to 1981-2000 69 Aus: Regional warming may have led to an increase in hydrological drought (low confidence)62,63 NZ: Mean westerly flow to increase by around 20% in spring and around 70% in winter, and to decrease by around 20% in summer and autumn, by 20905 NZ: Extreme westerlies and southerlies have slightly increased while extreme easterlies have decreased since 196013,71 Aus: Off shore regional sea level rise may exceed 10% more than global SLR, see AR5 WGI Chap13, Figure 13.2174 NZ: Off shore regional sea level rise may be up to 10% more than global SLR75 Aus and NZ: Satellite estimates of regional SLR for 1993-2009 are significantly higher than those for 1920-2000, partly reflecting climatic variability72,73,76,77 Aus: An increase of mean sea level by 0.1m increases the frequency of an extreme sea level event by a factor of between 2 and 10 over southeastern Australia depending on location 80-82, Aus: Increase in days with very high and extreme fire danger index by 2-30% (2020), 5100% (2050) (using B1 and A2 and two climate models, and 1973-2007 base)85 NZ: Increase in days with very high and extreme fire danger index from around 0 to 400% (2040) and 0 to 700% (2090) (using A1B,16 CMIP3 GCMs )89 Aus: Modelling study shows a 50% reduction in TC occurrence for 2051-2090 relative to 1971-2000, increases in intensity of the modelled storms, and occur around 100km further south94 NZ: Occurrence of conditions conducive to convective storm development is projected to increase by 3–6% by 2070-2100 (A2), relative to 1970-2000, with the largest increases over the South Island70 Aus and NZ: Many of past and projected changes in mean wind speed can be related to changes in atmospheric circulation43,67,68 NZ: Allowing for glacial isostatic adjustment, absolute observed SLR is around 2.0mm/yr73,78 Aus: For the example of Canberra, the projected changes represent the current 17 days per year increasing to 18-23 days in 2020 and 20-33 days in 205085 Aus: Regional research on convective storms is limited but studies have shown a projected decrease in the frequency of cool-season tornadoes95, and hail3 in southern Australia, and increases in the frequency and intensity of hail in the Sydney region3,96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Snow and ice Aus: Late season significant snow depth decline at three out of four Snowy mountain sites over 1957-200297 (**) NZ: Ice volume declined by 36-61% from the mid-late 1800s to the late 1900s98-100, with glacier volume reducing by 15% between 1976 and 2008101 (**) Aus: Both snow depth and area are projected to decline97 (***) NZ: Snowline elevations are projected to rise, and winter snow volume and days with low elevation snow cover are projected to decrease5,102,103 (***) Aus: Area with at least 30 days cover annually projected to decline 14-54% (2020) and 3093% (2050)97 NZ: By 2090, peak snow accumulation is projected to decline by 32-79% at 1000m and by 6-51% at 2000 m103 NZ: Atmospheric circulation variations can enhance or outweigh multi-decadal trends in ice volume over time scales of up to two decades104,105 References: 1 Fawcett et al. (2012); 2 Mullan et al. (2010); 3 CSIRO and BoM (2007); 4 Moise and Hudson (2008); 5 MfE (2008b); 6 AR5-WGI-Atlas-AI68-69; 7 AR5-WGI-Ch11; 8 AR5-WGI-Ch12; 9 AR5-WGI-Ch14; 10 Karoly and Braganza (2005); 11 Hendon et al. (2007); 12 Nicholls et al. (2010); 13 Dean and Stott (2009); 14 Lough (2008); 15 Lough and Hobday (2011); 16 Chambers and Griffiths (2008); 17 Gallant and Karoly (2010); 18 Nicholls and Collins (2006); 19 Trewin and Vermont (2010); 20 BOM (2013); 21 Alexander and Arblaster (2009); 22 Tryhorn and Risbey (2006); 23 Griffiths et al. (2005); 24 Tait (2008); 25 Alexander et al. (2007); 26 Deo et al. (2009); 27 McAlpine et al. (2007); 28 Cruz et al. (2010); 29 Hope et al. (2010); 30 Jones et al. (2009); 31 Gallant et al. (2012); 32 Griffiths (2007); 33 Timbal and Jones (2008); 34 AR5-WGI-Atlas-AI70-71; 35 Irving et al. (2012); 36 Watterson (2012); 37 Reisinger et al. (2010); 38 Bates et al. (2008); 39 Frederiksen and Frederiksen (2007); 40 Hope et al. (2006); 41 Timbal et al. (2006); 42 Cai and Cowan (2006); 43 Frederiksen et al. (2011); 44 Cai et al. (2011); 45 Nicholls (2010); 46 Smith and Timbal (2010); 47 Timbal et al. (2010a); 48 Gallant et al. (2007); 49 Westra and Sisson (2011); 50 Jakob et al. (2011); 51 Abbs and Rafter (2009); 52 Rafter and Abbs (2009); 53 Kharin et al. (2013); 54 IPCC-SREX-Chapter-3; 55 Westra et al. (2013); 56 CareySmith et al. (2010); 57 Hennessy et al. (2008a); 58 Mullan et al. (2005); 59 Kirono and Kent (2010); 60 Kirono et al. (2011); 61 Clark et al. (2011); 62 Cai and Cowan (2008); 63 Nicholls (2006); 64 Alexander et al. (2011); 65 McVicar et al. (2008); 66 Troccoli et al. (2012); 67 Parker et al. (2007); 68 Mullan et al. (2001); 69 McInnes et al. (2011a); 70 Mullan et al. (2011); 71 Salinger et al. (2005); 72 Burgette et al. (2013); 73 Hannah and Bell (2012); 74 AR5-WGI-Ch13; 75 Ackerley et al. (2013); 76 CSIRO and BoM (2012); 77 Meyssignac and Cazenave (2012); 78 Hannah (2004); 79 Menendez and Woodworth (2010); 80 McInnes et al. (2009); 81 McInnes et al. (2011b); 82 McInnes et al. (2012); 83 Harper et al. (2009); 84 Clarke et al. (2012); 85 Lucas et al. (2007); 86 Hasson et al. (2009); 87 Cai et al. (2009a); 88 Clarke et al. (2011); 89 Pearce et al. (2011); 90 Kuleshov et al. (2010); 91 Callaghan and Power (2011); 92 Hassim and Walsh (2008); 93 Allen and Karoly (2013); 94 Abbs (2012); 95 Timbal et al. (2010b); 96 Leslie et al. (2008); 97 Hennessy et al. (2008b); 98 Hoelzle et al. (2007); 99 Ruddell (1995); 100 Chinn (2001); 101 Chinn et al. (2012); 102 Fitzharris (2004); 103 Hendrikx et al. (2012); 104 Purdie et al. (2011); 105 Willsman et al. (2010). Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-2: Constraints and enabling factors for institutional adaptation processes in Australasia.* Constraint Uncertainty of projections Availability and cost of data and models Limited financial and human capability and capacity; time lag in developing expertise Unclear problem definition and goals; unclear standards for risk assessment methodologies and decision support tools; limited monitoring and evaluation Unclear or contradictory legislative frameworks and responsibilities, unclear liabilities Enabling factors Improved 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 Central provision of relevant core climate and non-climate data, including regional scenarios of projected changes 4,5,7,9,18,19 National first-pass risk assessments 4,5,7,8,18,20-24 Support for pilot projects 4,8,15,18,24,25 Building 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 Explicit but iterative framing and scoping of adaptation challenge, to reflect alternative entry points for stakeholders while meeting expectations of project sponsors to ensure long-term support 3,11,17,31-34 Tailoring decision-making frameworks to specific problems 1,2,6,17,35,36 Criteria and tools to monitor and evaluate adaptation success 7,18,37-39 Clear and coordinated legislative frameworks 5,8,9,15,24,40-45 Defined 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 of policies; disciplinary voids or single approaches Whole-of-council approach to climate adaptation to break up institutional and professional silos 15,33,47 Long-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 flexible land-use decisions 4,5,9,49 Strengthening multi-disciplinarity across professional fields 14,29,48 Lack of political leadership; Legally binding guidance and clarification of liabilities and duty of care to reduce short election cycles; limited dependence on individual leadership 5,7-9,15,24,38,40,46,49 community support, Consistent but audience-specific communication of current and potential future participation and awareness vulnerability and implications for community values 4,5,7,26,42,43,50 for adaptation Comprehensible communication of and access to response options, and their consistency with wider development plans 7,26,28,33,39,42,43 Clearly identified entry points for public participation 17,34,38,39,42,48,51-53 * Note: The relevance of each constraint varies among organisations, sectors and location. 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. (2012); 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). Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-3: Examples of detected changes in species, natural and managed ecosystems, consistent with a climate change1 signal, published since the AR4. Confidence 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. Confidence 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 insufficient to explain the observed change. Type of change and nature of evidence Morphology Limited evidence (1 study) Geographic distribution High agreement, robust evidence for many marine species & mobile terrestrial species Examples Declining body size of southeast Australian passerine birds, equivalent to ~7o latitudinal shift (Gardner et al., 2009) Southerly range extension of the barrens-forming sea urchin Centrostephanus rodgersii from the New South Wales coast to Tasmania; flow on impacts to marine communities including lobster fishery; shift of 160 km per decade over 30 years (Ling, 2008; Ling et al., 2008; Ling et al., 2009; Banks et al., 2010) Forty-five fish species, representing 27 families (about 30% of the inshore fish families occurring in the region), exhibited major distributional shifts in Tasmania (Last et al., 2011) 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) Subject to Final Copyedit Time scale of observations Confidence in the detection of biological change ~100 years medium trend significant for 4 out of 8 species, two other species show same trend but not statistically significant high ~30-50 years (first recorded in Tasmania late 1970s) distributions from late 1880s, 1980s and present (1995now) ~50 years Sites resampled 2007-2008, compared with 1950s 89 Potential climate change driver(s) 1 Warming air temperatures ~1.0oC over same period Confidence in the role of climate vs other drivers medium Nutritional cause discounted Increased sea surface temperature (SST), Ocean warming in SE Australia, increased southerly penetration of the East Australian Current (EAC), 350 km over 60 years high high Increased SST SE Australia, increased southerly penetration of EAC medium Changed fishing practices have potentially contributed to trends medium Increased SST in SE Australia (average 0.22oC per decade), increased southerly penetration of the EAC, 350 km over 60 years medium 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Life cycles Marine productivity Limited evidence, medium agreement 65 years high Increase in local air temperatures of 0.16oC per decade (1945-2007) 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 datasets (Chambers et al., 2013b) Multiple time periods from 1960s, all included 1990s and 2000s high Earlier wine-grape ripening at 9 of 10 sites in Australia (Webb et al., 2012) Multiple time periods up to 64 years (average 41 years) high Local climate trends (increasing air temperature, decreased raindays) were more important than broad-scale drivers such as the Southern Oscillation Index. Strongest associations were with decreased raindays. Increased length of growing season, increased average temperature and reduced soil moisture Timing of migration of glass eels, Anguilla spp. advanced by several weeks in Waikato River, North island, New Zealand (Jellyman et al., 2009) Robust evidence, medium agreement; increasing documentation of advances in phenology in some species (mainly migration and reproduction in birds, emergence in butterflies, flowering in plants) but also significant trends towards later life cycle events in some taxa (see meta-analysis for Southern Hemisphere phenology (Chambers et al., 2013a) Significant advance in mean emergence date of 1.5 days per decade (1941-2005) in the Common Brown butterfly Heteronympha merope in Australia (Kearney et al., 2010) 30 years (2004-2005 compared to 1970s) Birth years ranged 18611993 (fish 2-128 years old) medium Warming water temperatures in spawning grounds high 60 year dataset; decline recorded over period 1997-2007 high Increasing growth rates in species in top 250m associated with warming SST, declining growth rates in species >1000m associated with long-term cooling (as indicated by Mg/Ca ratios and delta18O in deep water corals) Increased SST and extension EAC associated with reduced nutrient availability Otolith (“ear stone”) analyses in long-lived Pacific fish indicates significantly increased growth rates 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 observed in the 2 intermediate-depth species (Thresher et al., 2007) ~50% decline in growth rate and biomass of spring phytoplankton bloom in western Tasman Sea (Thompson et al., 2009) Subject to Final Copyedit 90 high Advance consistent with physiologically based model of temperature influence on development high No other potential confounding facotrs identified medium Changed husbandry techniques, resulting in lower crop yields, may have contributed to trend low Changes in discharge discounted as contributing factor medium Changed fishing pressure may have contributed to trend medium 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Vegetation change Limited agreement & evidence; interacting impacts of changed land practices, altered fire regimes, increasing atmospheric CO2 concentration and climate trends difficult to disentangle Freshwater communities Limited evidence (1 study) Disease Limited evidence, robust agreement Coral reefs Robust evidence & high agreement Expansion of monsoon rainforest at expense of eucalypt savanna and grassland in Northern Territory, Australia (Banfai and Bowman, 2007; Bowman et al., 2010) ~40 years medium Increases in rainfall and atmospheric CO2 Net increase in mire wetland extent (10.2%) and corresponding contraction of adjacent eucalypt woodland in seven sub-catchments in south east Australia (Keith et al., 2010) Weather data covers >40 years (depending on parameter); vegetation mapping from 1961-1998 13 years (1994-2007) medium Decline in evapo-transpiration medium Increasing water temperatures and declining flows 1998 onwards medium Increasing SST low Variation in sampling, changes in water quality, impacts of impoundment and water extraction may have contributed to trends high 1979 onwards high Increasing SST high 1971-2003; 1990-2005 high Increasing SST high Changes in water quality discounted Decline in families of macroinvertebrates that favour cooler, faster-flowing habitats in New South Wales streams and increase in families favouring warmer and more lentic conditions (Chessman, 2009) Emergence and increased incidence of coral diseases including white syndrome (since 1998), and black band disease (since 1993-4) (Bruno et al., 2007; Sato et al., 2009; Dalton et al., 2010) Multiple mass bleaching events since 1979 (see 25.6.2, 30.5) Calcification 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) Subject to Final Copyedit 91 medium Changes in fire regimes and land management practices may have contributed to trend low Resource exploitation, fire history and autogenic mire development discounted 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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). 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 rutaceaous species and endemic psyllid fauna altered distribution of existing invasive & pathogenic species Nassella neesiana (Chilean needle grass): increased droughts favour establishment 2 Warming and drying may encourage the spread of existing invasives such as Pheidole megacephala in New Zealand and provide suitable conditions for other exotic ant species if they invade 3 Reduced climatic suitability for exotic invasive grasses in Australia (11 species including Nassella sp.) 4 Range of the invasive weed Lantana camara (lantana) projected to extend from Northern Australia to Victoria, South Australia and Tasmania 5 Projected increases in the range of three recently naturalised sub-tropical plants (Archontophoenix cunninghamiana, Psidium guajava, Schefflera actinophylla) 6 Managed pasture in New Zealand Human health and potentially agricultural and natural ecosystems Australian rangeland Multiple Native ecosystems in New Zealand altered climatic constraints on invasive & pathogenic species Queensland fruit fly (Bactrocera tryoni) moving southwards 7 Significant 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 Australian horticulture Native frogs altered impact of existing invasive & pathogenic species Fusarium pseudograminearum causing crown rot increases under elevated CO2 9 Increased abundance of the root-feeding nematode Longidorus elongatus under elevated CO2 10 Australian wheat New Zealand pasture Increased severity of Swiss needle cast disease caused by Phaeocryptopus gaeumannii 11 Douglas fir plantations in New Zealand, impact more severe in North Island Light brown apple moth, Epiphyas postvittana (Walker) (Lepidoptera:Tortricidae) reduction in natural enemies due to asynchrony and loss of host species 12 Projected changes in the efficacy of five biological control systems demonstrating a range of potential disruption mechanisms 13 Australian horticulture altered effectiveness of management strategies Pastoral and horticultural systems in New Zealand References: 1 Finlay et al. (2009); 2 Bourdôt et al. (2012); 3 Harris and Barker (2007); 4 Gallagher et al. (2012a); 5 Taylor et al. (2012b); 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). Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-5: Examples of co-beneficial 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. Climate impact Hot days and heatwaves 1-8 Adaptation options Greening cities/roofs; more green spaces; well-designed energy efficient buildings; occupant behavioral change; standards for new and retrofitting of existing infrastructure and assets; new methods and material for transport infrastructure to withstand higher extreme temperature Co-benefits Energy efficiency; reduced risk of blackouts; fewer health impacts; resilient infrastructure and assets; resilient community Barriers to adoption Lack of standards; high installation costs; limited understanding of benefits; high individual discount rate; split of private costs and public benefits Decreased water supply and drought Supply augmentation (water recycling, rainwater harvesting, increased storage, desalinisation); demand management; infrastructure upgrades; integrated urban sensitive design Water self-sufficiency for current and future demand/population; less pipe/storage leakage; reduced environmental impacts from abstraction New standards and improvements to building, water infrastructure (e.g. drainage and sewerage) and transport infrastructure; upgrades of protection systems; retaining floodplains/floodways; restoring wetlands; buffers from hazardprone areas; raising minimum floor levels; rezoning/ relocation New building design to withstand higher wind pressures; rezoning/relocation Reduced damages to homes and infrastructure and loss of life; decreased insurance premiums; habitat protection Potential health impacts of recycled water; lower than expected uptake of demand options and relaxation after crises; trade-offs between supply and demand management; cost and environmental impacts of some augmentation options High implementation cost especially if retrospective on existing stock; rezoning/ relocation can affect property prices and are highly contested [See Box 25-2 for more] River and local flooding, coastal erosion and inundation [See Boxes 25-1 and 25-8 for more] Severe storms and tropical cyclones 9-12 Corrosion from increased atmospheric CO2 levels 13,14 Improved standards for construction using concrete; application of coatings for existing building stock Reduced damages to homes and infrastructure and loss of life; decreased insurance premiums Reduced rates of carbonation-induced corrosion of concrete High implementation cost; rezoning/ relocation can affect property prices and are highly contested Effectiveness of coatings varies with age and condition of concrete 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); 10 Wang et al. (2010b); 11 Stewart and Wang (2011); 12 Mason et al. (2013); 13 Stewart et al. (2012); 14 Wang et al. (2012). Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 conflict (cause negative impacts) or be synergistic (have co-benefits) 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 impacts and adaptation responses Reduction of bushfire risk in natural landscapes Biodiversity, tourism Potential for greater conflict between conservation managers and other park users in Kosciuszko National Park if increasing fire incidence causes park closures, either to reduce risk, or to rehabilitate vegetation after fires (Wyborn, 2009), e.g. Objectives of the Wildfire Management Overlay (WMO) in Victoria conflicts with vegetation conservation (Hughes and Mercer, 2009). Reduction of risk to energy transmission from bushfires Biodiversity, energy Underground cabling would reduce both the susceptibility of transmission networks to fire and ignition sources for wild fires, thus reducing risks to ecosystems and settlements; constraints include significant investment cost, diverse ownership of assets and lack of an overarching national strategy (ATSE, 2008; Parsons Brinkerhoff, 2009; Linnenluecke et al., 2011). Protection of coastal infrastructure Biodiversity, tourism Seawalls may provide habitat but these communities have different diversity and structure to those developing on natural substrates (Jackson et al., 2008); groynes potentially alter beach fauna diversity and community structure (Walker et al., 2008); continuing hard protection against sea level rise results in long-term loss of coastal amenities (Gorddard et al., 2012). Avoidance of risks from sea level rise via relocation Indigenous 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 et al., 2010b; McNamara et al., 2011). Allocating scarce water resources via market instruments Rural areas, agriculture, mining Market 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 water supplies, with potentially significant social, environmental and wider economic consequences (Kiem and Austin, 2012). Increased water security via augmentation of supply for urban and agricultural systems Biodiversity, Water storage can buffer urban settlements and agricultural systems against high variability in river flows, but altered flow regimes can have significant negative impacts on freshwater ecosystems (Bond et al., 2008; Pittock et al., 2008; Kingsford, 2011). Discharge 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); increasing supply can reduce the effectiveness of demand-side measures (Barnett and O'Neill, 2010; Taptiklis, 2011; Box 25-2). Subject to Final Copyedit water demand management 94 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-7: Examples of interactions between adaptation and mitigation measures (green rows denote synergies where multiple benefits may be realized, orange rows denote potential tradeoffs and conflicts; grey row gives an example of complex, mixed interactions). The primary goal may be adaptation or mitigation. 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 2004-05. 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 efficiency (Wang et al., 2010a) and increase energy demand and associated CO2 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 localised 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 modelled (Smales, 2006). Urban densification 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 flood 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 CO2 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 towards perennial C4 grasses would increase methane emissions from grazing ruminants due to lower feed quality, but studies in south-west 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 benefits 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 benefits, 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 fires Biodiversity, livelihoods Improved management of savanna fires to reduce the extent of high intensity late season fires could substantially reduce emissions as well as having significant benefits 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 significant biodiversity benefits (NRMMC, 2010; Bradshaw et al., 2013). Economic benefits of reduced grazing competition, infrastructure damage and greenhouse gases could outweigh costs of camel reductions (Drucker et al., 2010). Subject to Final Copyedit 95 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 20802100), 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. Subject to Final Copyedit 96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-1: Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression [WGI AR5 Figures SPM.1 and 2.21]. (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression [WGI AR5 Figure SPM.2]. For observed temperature and precipitation, trends have been calculated where sufficient data permits 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 change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, 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, less than the baseline variability in >66% of models. (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 21-3 and CC-RC]. Subject to Final Copyedit 97 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-2: Observed and simulated variations in past and projected future annual average near-surface air temperature over land areas of Australia (left) and New Zealand (right). Black lines show various estimates from observational measurements. Shading denotes the 5-95 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 ‘RCP2.6’ emissions scenario (63), and the ‘RCP8.5’ (63). Data are anomalies from the 1986-2005 average of the individual observational data (for the observational time series) or of the corresponding historical allforcing simulations. Further details are given in Box 21-3. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 98 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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. Subject to Final Copyedit 99 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 (millimetres; 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). Subject to Final Copyedit 100 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 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 ~1990 (based on available meteorological station data for the period 1975-2004), ~2050 and ~2100. Bar charts show the change in population heat exposure, expressed as persondays 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 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). Subject to Final Copyedit 101 28 October 2013