FINAL DRAFT                       Chapter 11               IPCC WGII Sixth Assessment Report

1

2                                     Chapter 11: Australasia

3

4 Coordinating Lead Authors: Judy Lawrence (New Zealand) and Brendan Mackey (Australia)

5

6 Lead Authors: Francis Chiew (Australia), Mark J. Costello (New Zealand/Norway/Ireland), Kevin
7 Hennessy (Australia), Nina Lansbury (Australia), Uday Bhaskar Nidumolu (Australia), Gretta Pecl
8 (Australia), Lauren Rickards (Australia), Nigel Tapper (Australia), Alistair Woodward (New Zealand), Anita
9 Wreford (New Zealand)

10

11 Contributing Authors: Jason Alexandra (Australia), Anne-Gaelle Ausseil (New Zealand), Shaun Awatere
12 (New Zealand), Douglas Bardsley (Australia), Rob Bell (New Zealand), Paula Blackett (New Zealand),
13 Sarah Boulter (Australia), Nicholas Cradock-Henry (New Zealand), Sandra Creamer (Australia), Rebecca
14 Darbyshire (Australia), Sam Dean (New Zealand), Alejandro Di Luca (Australia/Canada), Andrew Dowdy
15 (Australia), Joanna Fountain (New Zealand), Michael Grose (Australia), Stefan Hajkowicz (Australia),
16 David Hall (New Zealand), Sarah Harris (Australia), Peter Hayman (Australia), Jane Hodgkinson
17 (Australia), Karen Hussey (Australia), Rhys Jones (New Zealand), Darren King (New Zealand), Martina
18 Linnenluecke (Australia), Erich Livengood (New Zealand), Mary Livingston (New Zealand), Cate MacInnis-
19 Ng (New Zealand), Belinda McFadgen (New Zealand), Celia McMichael (Australia), Bradley Moggridge
20 (Australia), Sandy Morrison (New Zealand), Vinnitta Mosby (Australia), Esther Onyango (Australia/Kenya),
21 Sharanjit Paddam (Australia), Grant Pearce (New Zealand), Petra Pearce (New Zealand), Rosh Ranasinghe
22 (Netherlands), David Schoeman (Australia), Rodger Tomlinson (Australia), Michael Watt (New Zealand),
23 Seth Westra (Australia), Russell Wise (Australia).

24

25 Review Editors: Ove Hoegh-Guldberg (Australia), David Wratt (New Zealand)

26

27 Chapter Scientists: Belinda McFadgen (New Zealand), Esther Onyango (Australia/Kenya)

28

29 Date of Draft: 1 October 2021

30

31 Notes: TSU Compiled Version

32

33

34 Table of Contents

35

36 Executive Summary..........................................................................................................................................3

37 11.1 Introduction ..............................................................................................................................................7

38  11.1.1 Context ............................................................................................................................................7

39  11.1.2 Economic, Demographic and Social Trends...................................................................................9

40 11.2 Observed and Projected Climate Change ..............................................................................................9

41  11.2.1 Observed Climate Change ..............................................................................................................9

42  11.2.2 Projected Climate Change ............................................................................................................14

43 11.3 Observed Impacts, Projected Impacts and Adaptation ......................................................................19

44  11.3.1 Terrestrial and Freshwater Ecosystems .......................................................................................19

45 Box 11.1: Escalating Impacts and Risks of Wildfire ...................................................................................27

46  11.3.2 Coastal and Ocean Ecosystems ....................................................................................................31

47 Box 11.2: The Great Barrier Reef in Crisis..................................................................................................37

48  11.3.3 Freshwater Resources...................................................................................................................40

49 Box 11.3: Drought, Climate Change, and Water Reform in the Murray-Darling Basin ........................44

50 Box 11.4: Changing Flood Risk .....................................................................................................................45

51  11.3.4 Food, Fibre, Ecosystem Products .................................................................................................46

52 Box 11.5: New Zealand's Land, Water and People Nexus under a changing climate .............................51

53  11.3.5 Cities, Settlements and Infrastructure...........................................................................................52

54 Box 11.6: Rising to the Sea-Level Challenge ................................................................................................57

55  11.3.6 Health and Wellbeing....................................................................................................................60

56  11.3.7 Tourism .........................................................................................................................................63

57  11.3.8 Finance..........................................................................................................................................64

    Do Not Cite, Quote or Distribute  11-1                                               Total pages: 151
    FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1   11.3.9 Mining ...........................................................................................................................................66

2   11.3.10 Energy ...................................................................................................................................66

3   11.3.11 Detection and Attribution of Observed Climate Impacts......................................................68

4 11.4 Indigenous Peoples .................................................................................................................................69

5   11.4.1 Aboriginal and Torres Strait Islander Peoples of Australia.........................................................69

6   11.4.2 Tangata Whenua ­ New Zealand Mori .......................................................................................72

7 11.5 Cross-Sectoral and Cross-Regional Implications ................................................................................73

8   11.5.1 Cascading, compounding and aggregate impacts ........................................................................73

9   11.5.2 Implications for National Economies ...........................................................................................75

10 11.6 Key Risks and Benefits...........................................................................................................................77

11 11.7 Enabling Adaptation Decision-making.................................................................................................85

12  11.7.1 Observed Adaptation Decision-Making.......................................................................................85

13  11.7.2 Barriers and Limits to Adaptation ................................................................................................90

14  11.7.3 Adaptation enablers ......................................................................................................................92

15 11.8 Climate Resilient Development Pathways............................................................................................97

16  11.8.1 System Adaptations and Transitions .............................................................................................97

17  11.8.2 Challenges for Climate Resilient Development Pathways............................................................98

18 FAQ 11.1: How is climate change affecting Australia and New Zealand? ...............................................98

19 FAQ 11.2: What systems in Australia and New Zealand are most at risk from ongoing climate change?

20  .................................................................................................................................................................. 99

21 FAQ 11.3: How can Indigenous Peoples' knowledge and practice help us understand contemporary

22  climate impacts and inform adaptation in Australia and New Zealand? .......................................100

23 FAQ 11.4: How can Australia and New Zealand adapt to climate change?...........................................100

24 References......................................................................................................................................................102

25

26

27

    Do Not Cite, Quote or Distribute  11-2        Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Executive Summary

 2

 3 Observed changes and impacts

 4

 5 Ongoing climate trends have exacerbated many extreme events (very high confidence). The Australian
 6 trends include further warming and sea-level rise, with more hot days and heatwaves, less snow, more
 7 rainfall in the north, less April-October rainfall in the south-west and south-east, more extreme fire weather
 8 days in the south and east. The New Zealand trends include further warming and sea-level rise, more hot
 9 days and heatwaves, less snow, more rainfall in the south, less rainfall in the north, and more extreme fire
10 weather in the east. There have been fewer tropical cyclones and cold days in the region. Extreme events
11 include Australia's hottest and driest year in 2019 with a record-breaking number of days over 39oC, New
12 Zealand's hottest year in 2016, three widespread marine heatwaves during 2016-2020, Category 4 cyclone
13 Debbie in 2017, seven major hailstorms over eastern Australia and two over New Zealand from 2014-2020,
14 three major floods in eastern Australia and three over New Zealand during 2019-2021, and major fires in
15 southern and eastern Australia during 2019-2020. {11.2.1, Table 11.2, 11.3.8}

16

17 Climate trends and extreme events have combined with exposure and vulnerabilities to cause major
18 impacts for many natural systems, with some experiencing or at risk of irreversible change in
19 Australia (very high confidence) and in New Zealand (high confidence). For example, warmer conditions
20 with more heatwaves, droughts and catastrophic wildfires have negatively impacted terrestrial and
21 freshwater ecosystems. The Bramble Cay melomys, an endemic mammal species, became extinct due to loss
22 of habitat associated with sea-level rise and storm surges in the Torres Strait. Marine species abundance and
23 distributions have shifted polewards, and extensive coral bleaching events and loss of temperate kelp forests
24 have occurred, due to ocean warming and marine heatwaves across the region. In New Zealand's Southern
25 Alps, from 1978-2016, the area of 14 glaciers declined 21%, and extreme glacier mass loss was at least six
26 times more likely in 2011, and ten times more likely in 2018, due to climate change. The end-of-summer
27 snowline elevation for 50 glaciers rose 300m from 1949-2019. {11.3.1.1, 11.3.2.1, Table 11.2b, Table 11.4,
28 Table 11.6, Table 11.9}

29

30 Climate trends and extreme events have combined with exposure and vulnerabilities to cause major
31 impacts for some human systems (high confidence). Socio-economic costs arising from climate variability
32 and change have increased. Extreme heat has led to excess deaths and increased rates of many illnesses.
33 Nuisance and extreme coastal flooding have increased due to sea-level rise superimposed upon high tides
34 and storm surges in low-lying coastal and estuarine locations, including impacts on cultural sites, traditions
35 and lifestyles of Aboriginal and Torres Strait Islander Peoples in Australia and Tangata Whenua Mori in
36 New Zealand. Droughts have caused financial and emotional stress in farm households and rural
37 communities. Tourism has been negatively affected by coral bleaching, fires, poor ski seasons and receding
38 glaciers. Governments, business and communities have experienced major costs associated with extreme
39 weather, droughts and sea-level rise. {11.3, 11.4, 11.5.2, Table 11.2, Boxes 11.1-11.6}

40

41 Climate impacts are cascading and compounding across sectors and socio-economic and natural
42 systems (high confidence). Complex connections are generating new types of risks, exacerbating existing
43 stressors and constraining adaptation options. An example is the impacts that cascade between
44 interdependent systems and infrastructure in cities and settlements. Another example is the 2019-2020 south-
45 eastern Australian wildfires which burned 5.8 to 8.1 million hectares, with 114 listed threatened species
46 losing at least half of their habitat and 49 losing over 80%, over 3,000 houses destroyed, 33 people killed, a
47 further 429 deaths and 3230 hospitalizations due to cardiovascular or respiratory conditions, $1.95 billion in
48 health costs, $2.3 billion in insured losses, and $3.6 billion in losses for tourism, hospitality, agriculture and
49 forestry. {11.5.1, Box 11.1}

50

51 Increasing climate risks are projected to exacerbate existing vulnerabilities and social inequalities and
52 inequities (high confidence). These include inequalities between Indigenous and non-Indigenous Peoples,
53 and between generations, rural and urban areas, incomes and health status, increasing the climate risks and
54 adaptation challenges faced by some groups and places. Resultant climate change impacts include the
55 displacement of some people and businesses, and threaten social cohesion and community wellbeing.
56 {11.3.5, 11.3.6, 11.3.10, 11.4}

57

Do Not Cite, Quote or Distribute  11-3        Total pages: 151
    FINAL DRAFT                           Chapter 11                    IPCC WGII Sixth Assessment Report

1 Projected impacts and key risks

2

3 Further climate change is inevitable, with the rate and magnitude largely dependent on the emission
4 pathway (very high confidence1). Ongoing warming is projected, with more hot days and fewer cold days
5 (very high confidence). Further sea-level rise, ocean warming and ocean acidification are projected (very
6 high confidence). Less winter and spring rainfall is projected in southern Australia, with more winter rainfall

7 in Tasmania, less autumn rainfall in south-western Victoria and less summer rainfall in western Tasmania
8 (medium confidence), with uncertain rainfall changes in northern Australia. In New Zealand, more winter

 9 and spring rainfall is projected in the west and less in the east and north, with more summer rainfall in the
10 east and less in the west and central North Island (medium confidence). In New Zealand, ongoing significant
11 clean ice glacier retreat is projected (very high confidence). More droughts and extreme fire weather are

12 projected in southern and eastern Australia (high confidence) and over most of New Zealand (medium
13 confidence). Increased rainfall intensity is projected, with fewer tropical cyclones and a greater proportion of

14 severe cyclones (medium confidence). {11.2.2, Table 11.3, Box 11.6}

15

16 Climate risks are projected to increase for a wide range of systems, sectors and communities, which
17 are exacerbated by underlying vulnerabilities and exposures (high confidence) {11.3; 11.4}. Nine key

18 risks were identified, based on magnitude, likelihood, timing and adaptive capacity {11.6, Table 11.14}:

19

20 Ecosystems at critical thresholds, where recent climate change has caused significant damage and further

21 climate change may cause irreversible damage, with limited scope for adaptation

22

23  Loss and degradation of coral reefs and associated biodiversity and ecosystem service values in Australia

24  due to ocean warming and marine heatwaves, e.g. three marine heatwaves on the Great Barrier Reef

25  during 2016-2020 caused significant bleaching and loss. {11.3.2.1, 11.3.2.2, Box 11.2}

26

27  Loss of alpine biodiversity in Australia due to less snow, e.g. loss of alpine vegetation communities

28  (snow patch Feldmark and short alpine herb-fields) and increased stress on snow-dependent plant and

29  animal species, {11.3.1.1, 11.3.1.2}

30

31 Key risks that have potential to be severe but can be reduced substantially by rapid, large-scale and

32 effective mitigation and adaptation

33

34  Transition or collapse of alpine ash, snowgum woodland, pencil pine and northern jarrah forests in

35  southern Australia due to hotter and drier conditions with more fires, e.g. declining rainfall in southern

36  Australia over the past 30 years has led to drought-induced canopy dieback across a range of forest and

37  woodland types, and death of fire-sensitive tree species due to unprecedented wildfires. {11.3.1.1,

38  11.3.1.2}

39

40  Loss of kelp forests in southern Australia and southeast New Zealand due to ocean warming, marine

41  heatwaves and overgrazing by climate-driven range extensions of herbivore fish and urchins, e.g. less

42  than 10% of giant kelp in Tasmania was remaining by 2011 due to ocean warming. {11.3.2.1, 11.3.2.2}

43

44  Loss of natural and human systems in low-lying coastal areas due to sea-level rise, e.g. for 0.5 m sea-

45  level rise, the value of buildings in New Zealand exposed to 1-in-100 year coastal inundation could

46  increase by NZ$12.75 billion and the current 1-in-100 year flood in Australia could occur several times a

47  year. {11.3.5; Box 11.6}

48

49  Disruption and decline in agricultural production and increased stress in rural communities in south-

50  western, southern and eastern mainland Australia due to hotter and drier conditions, e.g. by 2050, a

51  decline in median wheat yields of up to 30% in south-west Australia and up to 15% in South Australia,

52  and increased heat stress in livestock by 31­42 days per year. {11.3.4; 11.3.5; Box 11.3}

    1 In this Report, the following summary terms are used to describe the available evidence: limited, medium, or robust;
    and for the degree of agreement: low, medium, or high. A level of confidence is expressed using five qualifiers: very
    low, low, medium, high, and very high, and typeset in italics, e.g., medium confidence. For a given evidence and
    agreement statement, different confidence levels can be assigned, but increasing levels of evidence and degrees of
    agreement are correlated with increasing confidence.

    Do Not Cite, Quote or Distribute      11-4                                                 Total pages: 151
    FINAL DRAFT                            Chapter 11         IPCC WGII Sixth Assessment Report

1   Increase in heat-related mortality and morbidity for people and wildlife in Australia due to heatwaves,

2   e.g. heat-related excess deaths in Melbourne, Sydney and Brisbane are projected to increase by about

3   300/year (low emission pathway) to 600/year (high emission pathway) during 2031-2080 relative to

4   142/year during 1971-2020. {11.3.1, 11.3.5.1, 11.3.5.2, 11.3.6.1, 11.3.6.2}

5

6 Key cross-sectoral and system-wide risk

7

8   Cascading, compounding and aggregate impacts on cities, settlements, infrastructure, supply-chains and

9   services due to wildfires, floods, droughts, heatwaves, storms and sea-level rise, e.g. in New Zealand,

10  extreme snow, heavy rainfall and wind events have combined to impact road networks, power and water

11  supply, interdependent wastewater and stormwater services and business activities {11.3.3, 11.5.1,

12  11.8.1}.

13

14 Key implementation risk

15

16  Inability of institutions and governance systems to manage climate risks, e.g. the scale and scope of

17  projected climate impacts overwhelm the capacity of institutions, organisations and systems to provide

18  necessary policies, services, resources and coordination to address the socio-economic impacts {11.5.1.2,

19  11.5.1.3, 11.5.2.3, 11.6, 11.7.1, 11.7.2, 11.7.3}.

20

21 There are important interactions between mitigation and adaptation policies and their implementation

22 (high confidence). Integrated policies in interdependent systems across biodiversity, water quality, water
23 availability, energy, transport, land use and forestry for mitigation, can support synergies between adaptation
24 and mitigation. These have co-benefits for the management of land use, water and associated conflicts, and

25 for the functioning of cities and settlements. For example, projected increases in fire, drought, pest

26 incursions, storms and wind place forests at risk and affect their ongoing role in meeting New Zealand's
27 emissions reduction goals. {11.3.4.3, 11.3.10.2, 11.3.5.3, Box 11.5}

28

29 Challenges and solutions

30

31 The ambition, scope and progress of the adaptation process has increased across governments, non-
32 government organisations, businesses and communities (high confidence). This process includes
33 vulnerability and risk assessments, identification of strategies and options, planning, implementation,

34 monitoring, evaluation and review. Initiatives include institutional frameworks in statute for risk assessment
35 and national adaptation planning and monitoring in New Zealand, a National Recovery and Resilience
36 Agency and National Disaster Risk Reduction Framework in Australia, deployment of new national
37 guidance, decision tools, collaborative governance approaches, and the introduction of climate risk and

38 disclosure regimes for the private sector. The focus however has been on adaptation planning, rather than on

39 implementation. {11.5.1, 11.7.1.1, Box 11.6, Table 11.15a, Table 11.15b, Table 11.17}

40

41 Adaptation progress is uneven, due to gaps, barriers and limits to adaptation, and adaptive capacity

42 deficits (very high confidence). Progress in adaptation planning, implementation, monitoring and evaluation

43 is lagging. Barriers include lack of consistent policy direction, competing objectives, divergent risk
44 perceptions and values, knowledge constraints, inconsistent information, fear of litigation, up-front costs,
45 and lack of engagement, trust and resources. Adaptation limits are being approached for some species and

46 ecosystems. Adaptive capacity to address the barriers and limits can be built through greater engagement
47 with groups and communities to build trust and social legitimacy through inclusion of diverse values,
48 including those of Aboriginal and Torres Strait Islander Peoples and Tangata Whenua Mori. {11.4, 11.5,
49 11.6, 11.7, 11.8, Table 11.4, Table 11.5, Table 11.6, Table 11.16, Box 11.2}

50

51 A range of incremental and transformative adaptation options and pathways is available as long as
52 enablers are in place to implement them (high confidence). Key enablers for effective adaptation include
53 shifting from reactive to anticipatory planning, integration and coordination across levels of government and
54 sectors, inclusive and collaborative institutional arrangements, government leadership, policy alignment,

55 nationally consistent and accessible information, decision-support tools, along with adaptation funding and

56 finance and robust consistent and strategic policy commitment. Over three-quarters of people in Australia

    Do Not Cite, Quote or Distribute                    11-5                              Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 and New Zealand agree that climate change is occurring and over 60% believe climate change is caused by
 2 humans, giving climate adaptation and mitigation action further social legitimacy. {11.7.3, Table 11.17}

 3

 4 New knowledge on system complexity, managing uncertainty and how to shift from reactive to
 5 adaptive implementation is critical for accelerating adaptation (high confidence). Priorities include:
 6 greater understanding of impacts on natural system dynamics; the exposure and vulnerability of different
 7 groups within society, including Indigenous Peoples; the relationship between mitigation and adaptation; the
 8 effectiveness and feasibility of different adaptation options; the social transitions needed for transformative
 9 adaptation; and the enablers for new knowledge to better inform decision making, e.g. monitoring data
10 repositories, risk and vulnerability assessments, robust planning approaches, sharing adaptation knowledge
11 and practice. {11.7.3.3}

12

13 Aboriginal and Torres Strait Islander Peoples and Tangata Whenua Mori can enhance effective
14 adaptation through the passing down of knowledge about climate change planning that promotes
15 collective action and mutual support across the region (high confidence). Supporting Aboriginal and
16 Torres Strait Islander Peoples and Tangata Whenua Mori institutions, knowledge and values, enables self-
17 determination and creates opportunities to develop adaptation responses to climate change. Actively
18 upholding the UN Declaration on the Rights of Indigenous Peoples and Mori interests under the Treaty of
19 Waitangi at all levels of government enables intergenerational approaches for effective adaptation. {11.3,
20 11.4, 11.6, 11.7.3; Cross-Chapter Box INDIG in Chapter 18}

21

22 A step change in adaptation is needed to match the rising risks and to support climate resilient
23 development (very high confidence). Current adaptation is largely incremental and reactive. A shift to
24 transformative and proactive adaptation can contribute to climate resilient development. The scale and scope
25 of cascading, compounding and aggregate impacts require new, larger-scale and timely adaptation.
26 Monitoring and evaluation of the effectiveness of adaptation progress and continual adjustment is critical.
27 The transition to climate-resilient development pathways can generate major co-benefits, but complex
28 interactions between objectives can create trade-offs. {11.7, 11.8.1, 11.8.2}

29

30 Delay in implementing adaptation and emission reductions will impede climate resilient development,
31 resulting in more costly climate impacts and greater scale of adjustments (very high confidence). The
32 region faces an extremely challenging future. Reducing the risks would require significant and rapid
33 emission reductions to keep global warming to 1.5-2.0oC, as well as robust and timely adaptation. The
34 projected warming under current global emissions reduction policies would leave many of the region's
35 human and natural systems at very high risk and beyond adaptation limits. {11.8, Table 11.1, Table 11.14,
36 Figure 11.6}

37

38

39

Do Not Cite, Quote or Distribute  11-6        Total pages: 151
    FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1 11.1 Introduction

2

3 This chapter assesses observed impacts, projected risks, vulnerability and adaptation, and the implications
4 for climate resilient development for the Australasia region, based on literature published up to 1 September
5 2021. It should be read in conjunction with other Working Group 2 chapters, the climate science assessment
6 in the Working Group 1 Report and the greenhouse gas emissions and mitigation assessment in the Working
7 Group 3 Report.

8

9 11.1.1 Context

10

11 The Australasia region is defined as the Exclusive Economic Zones (EEZ) and territories of Australia and
12 New Zealand. In both countries, climate adaptation is largely implemented at a sub-national level through
13 devolution of functions constitutionally or by statute, alongside disaster risk reduction (COAG, 2011;
14 Lawrence et al., 2015; Macintosh et al., 2015).

15

16 Australia's economy is dominated by financial and insurance services, education, mining, construction,
17 tourism, health care and social assistance (ABS, 2018) with Australian exports accruing mostly from mining
18 (ABS, 2018; ABS, 2019). In New Zealand, service industries, including tourism, collectively account for
19 around two thirds of GDP (NZ Treasury, 2016). The primary sector contributes 6% of New Zealand's GDP
20 and over half of the country's export earnings (NZ Treasury, 2016).

21

22 Existing vulnerabilities expose and exacerbate inequalities between rural, regional and urban areas,
23 Indigenous and non-Indigenous Peoples, those with health and disability needs, and between generations,
24 incomes and health status, increasing the relative climate change risk faced by some groups and places
25 (Jones et al., 2014; Bertram, 2015; Perry, 2017; Hazledine and Rashbrooke, 2018) (high confidence).

26

27 Previous IPCC reports (Table 11.1) have documented observed climate impacts, projected risks, adaptation
28 challenges and opportunities. In this chapter, there is more evidence of observed climate impacts and
29 adaptation, better quantification of socio-economic risks, new information about cascading and
30 compounding risks, greater emphasis on adaptation enablers and barriers, and links to climate-resilient
31 development.

32

33

34 Table 11.1: Summary of key conclusions from the IPCC 5th Assessment Report (AR5) Australasia chapter (Reisinger
35 et al., 2014) and relevant conclusions from the IPCC Special Reports on Global Warming of 1.5°C (IPCC, 2018),
36 Climate Change and Land (IPCC, 2019a) and Oceans and Cryosphere (IPCC, 2019b)

    Conclusions                                                                                             Report

    Our regional climate is changing (very high confidence) and warming will continue through the 21st      (Reisinger
    century (virtually certain) with more hot days, fewer cold days, less snow, less rainfall in southern   et al.,
    Australia and the northeast of both of New Zealand's islands, more rainfall in western New Zealand,     2014)
    more extreme rainfall, sea-level rise, increased fire weather in southern Australia and across New
    Zealand, and fewer cyclones but a greater proportion of intense cyclones.

    Key risks include changes in the structure and composition of Australian coral reefs, loss of montane
    ecosystems, increased flood damage, reduced water resources in southern Australia, more deaths and
    infrastructure damage during heatwaves, more fire-related impacts on ecosystems and settlements in
    southern Australia and across New Zealand, greater risk to coastal infrastructure and ecosystems, and
    reduced water availability in the Murray-Darling Basin and southern Australia (high confidence).
    Benefits are projected for some sectors and locations (high confidence), including reduced winter
    mortality and energy demand for heating, increased forest growth, and enhanced pasture productivity.

    Adaptation is occurring and adaptation is becoming mainstreamed in some planning processes (high
    confidence). Adaptive capacity is considered generally high in many human systems, but adaptation
    implementation faces major barriers, especially for transformational responses (high confidence). Some
    synergies and trade-offs exist between different adaptation responses, and between mitigation and
    adaptation, with interactions occurring both within and outside the region (very high confidence).

    Do Not Cite, Quote or Distribute  11-7                                                                 Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

Vulnerability remains uncertain due to incomplete consideration of socio-economic dimensions (very
high confidence), including governance, institutions, patterns of wealth and aging, access to technology
and information, labour force participation, and societal values.

Emissions reductions under Nationally Determined Contributions from signatories to the Paris                       (IPCC,
Agreement are consistent with a global warming of 2.5-3.0°C above pre-industrial temperatures by                   2018)
2100. Much deeper emission reductions are needed prior to 2030 to limit warming to 1.5°C. There are
limits to adaptation and adaptive capacity for some human and natural systems at global warming of
1.5°C, with associated losses.

Climate impacts will disproportionately affect the welfare of impoverished and vulnerable people                   (IPCC,
because they lack adaptation resources. Strengthening the climate-action capacities of national and sub-           2019a)
national authorities, civil society, the private sector, Indigenous people and local communities can
support implementation of actions.

Land-related responses that contribute to climate change adaptation and mitigation can also combat
desertification, land degradation, and enhance food security.

Appropriate design of policies, institutions and governance systems at all scales can contribute to land-
related adaptation and mitigation while facilitating the pursuit of climate-adaptive development
pathways.

Mutually supportive climate and land policies have the potential to save resources, amplify social
resilience, support ecological restoration, and foster collaboration between stakeholders.

Near-term action to address climate change adaptation and mitigation, desertification, land degradation
and food security can bring social, ecological, economic and development co-benefits. Delaying action
(both mitigation and adaptation) will be more costly.

The rate of global mean sea-level rise of 3.6 mm per year for 2006­2015 is unprecedented over the last             (IPCC,
century. Extreme wave heights, coastal erosion and flooding, have increased in the Southern Ocean by               2019b)
around 1.0 cm per year over the period 1985­2018.

Some species of plants and animals have increased in abundance, shifted their range, and established in
new areas as glaciers receded and the snow-free season lengthened. Some cold-adapted or snow-
dependent species have declined in abundance, increasing their risk of extinction, notably on mountain
summits.

Many marine species have shifted their range and seasonal activities. Altered interactions between
species have caused cascading impacts on ecosystem structure and functioning.

       Mean sea-level rise projections are higher by 0.1 m compared to AR5 under RCP8.5 in 2100. Extreme
       sea-level events that are historically rare (once per century) are projected to occur frequently (at least
       once per year) at many locations by 2050.

       Projected ecosystem responses include losses of species habitat and diversity, and degradation of
       ecosystem functions. Warm water corals are at high risk already and are projected to transition to very
       high risk even if global warming is limited to 1.5°C.

       Governance arrangements (e.g., marine protected areas, spatial plans and water management systems)
       are too fragmented across administrative boundaries and sectors to provide integrated responses to the
       increasing and cascading risks. Financial, technological, institutional and other barriers exist for
       implementing responses.

       Enabling climate resilience and sustainable development depends critically on urgent and ambitious
       emissions reductions coupled with coordinated, sustained and increasingly ambitious adaptation actions.
       This includes better cooperation and coordination among governing authorities, education and climate
       literacy, sharing of information and knowledge, finance, addressing social vulnerability and equity, and
       institutional support.

1
2

Do Not Cite, Quote or Distribute  11-8                                                               Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 11.1.2 Economic, Demographic and Social Trends

 2

 3 Economic, demographic and socio-cultural trends influence the exposure, vulnerability and adaptive capacity
 4 of individuals and communities (high confidence) (Elrick-Barr et al., 2016; Smith et al., 2016; Hayward,
 5 2017; B. Frame et al., 2018; Plummer et al., 2018; Smith et al., 2018; Gartin et al., 2020). In the absence of
 6 proactive adaptation, climate change, impacts are projected to worsen inequalities between Indigenous and
 7 non-Indigenous people and other vulnerable groups (Green et al., 2009; Manning et al., 2014; Ambrey et al.,
 8 2017) (high confidence). Socio-economic inequality, low incomes and high levels of debt, poor health and
 9 disabilities increase vulnerability and limit adaptation (Hayward, 2012) (11.7.2). Lack of services, such as
10 schools and medical services, in poorer and rural areas, and decision-making processes that privilege some
11 voices over others, exacerbate inequalities (Kearns et al., 2009; Hinkson and Vincent, 2018).

12

13 Changes to the composition and location of different demographic groups in the region contributes to
14 increased exposure or vulnerability to climate change (medium confidence). Australia's population reached
15 25 million in 2018 and is projected to grow to 37.4­49.2 million by 2066, with most growth in major cities
16 (accounting for 81% of Australia's population growth from 2016­17) (ABS, 2018), although COVID-19 is
17 expected to slow the growth rate (CoA, 2020c). Highest growth rates outside of major cities occurred mostly
18 in coastal regions (ABS, 2017) which have built assets exposed to sea-level rise. New Zealand's population
19 was 5.1 million at the end of 2020 and is projected to increase to 6.0­6.5 million by 2068 assuming no
20 marked changes in migration patterns (Stats NZ, 2016; Stats NZ, 2021). Although the population densities of
21 both countries are much lower than other OECD countries, they are highly urbanized with over 86% living in
22 urban areas in both countries (Productivity Commission, 2017; World Bank, 2018). This proportion is
23 projected to increase to over 90% by 2050 (UN DESA, 2019) mostly in coastal areas (Rouse et al., 2017).
24 Consideration of climate change impacts when planning and managing such growth and associated
25 infrastructure could help avoid new vulnerabilities being created, particularly from wildfires, sea-level rise,
26 heat stress and flooding.

27

28 The region has an increasingly diverse population through the arrival of migrants, including those from the
29 Pacific, whose innovations, skills and transnational networks enhance their and others' adaptive capacity (De
30 et al., 2016; Fatori et al., 2017; Barnett and McMichael, 2018), although language barriers and socio-
31 economic disadvantage can create vulnerabilities for some (11.7.2).

32

33 Climate change inaction exacerbates intergenerational inequity including prospects for the current younger
34 population (Hayward, 2012). Increasing transient worker populations (ABS, 2018) may diminish social
35 networks and adaptive capacity (Jiang et al., 2017). The region has an aging population and increasing
36 numbers of people living on their own who are highly vulnerable to extreme events, including heat stress and
37 flooding (Zhang et al., 2013).

38

39 Socio-economic trends are affected by global mega trends (KPMG, 2021), which are expected to influence
40 the region's ability to implement climate change adaptation strategies (World Economic Forum, 2014).
41 Digital technological advances have potential benefits for building adaptive capacity (Deloitte, 2017a).

42

43

44 11.2 Observed and Projected Climate Change

45

46 11.2.1 Observed Climate Change

47

48 Regional climate change has continued since the Fifth Assessment Report (AR5) in 2014, with trends
49 exacerbating many extreme events (very high confidence). The following changes are quantified with
50 references in Tables 11.2a and 11.2b. The region has continued to warm (Figure 11.1), with more extremely
51 high temperatures and fewer extremely low temperatures. Snow depths and glacier volumes have declined.
52 Sea-level rise and ocean acidification have continued. Northern Australia has become wetter, while April-
53 October rainfall has decreased in south-western and south-eastern Australia. In New Zealand, most of the
54 south has become wetter while most of the north has become drier (Figure 11.2). The frequency, severity and
55 duration of extreme fire weather conditions have increased in southern and eastern Australia and eastern
56 New Zealand. Changes in extreme rainfall are mixed. There has been a decline in tropical cyclone frequency
57 near Australia.

Do Not Cite, Quote or Distribute  11-9        Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1

 2 Reliable measurements are limited for some types of storms, particularly thunderstorms, lightning, tornadoes
 3 and hail (Walsh et al., 2016). Many high impact events are a combination of interacting physical processes
 4 across multiple spatial and temporal scales (e.g. fires, heatwaves and droughts), and better understanding of
 5 these extreme and compound events is needed (Zscheischler et al., 2018).

 6

 7 Some of the observed trends and events can be partly attributed to anthropogenic climate change, as
 8 documented in Chapter 16. Examples include regional warming trends and sea-level rise, terrestrial and
 9 marine heatwaves, declining rainfall and increasing fire weather in southern Australia, and extreme rainfall
10 and severe droughts in New Zealand.

11

12

13

14 Figure 11.1: Observed temperature changes in Australia and New Zealand. Annual temperature change time-series are
15 shown for 1910­2019. Mean annual temperature trend maps are shown for 1960­2019 using contours for Australia and
16 individual sites for New Zealand. Data courtesy of BoM and NIWA.

17

18

19

20 Figure 11.2: Observed rainfall changes in Australia and New Zealand. Rainfall change time-series for 1900­2019 are

21 shown for northern Australia (December-February: DJF), southwest Australia (June-August: JJA) and southeast

Do Not Cite, Quote or Distribute  11-10       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1 Australia (JJA). Dashed lines on the maps for Australia show regions used for the time-series. Rainfall trend maps are
2 shown for 1960­2019 (DJF and JJA) using contours for Australia and individual sites for New Zealand. Areas of low
3 Australian rainfall (less than 0.25 mm/day) are shaded white in JJA. Data courtesy of BoM and NIWA.

4

5

6 Table 11.2a: Observed climate change for Australia.

Climate variable Observed change                                                      References

Air temperature     Increased by 1.4°C from 1910­2019. 2019 was the warmest           (BoM, 2020b; BoM and
over land           year, and nine of the ten warmest on record occurred since        CSIRO, 2020; Trewin et al.,
                    2005. Clear anthropogenic attribution.                            2020; Gutiérrez et al., 2021)

Sea surface         Increased by 1.0°C from 1900-2019 (0.09oC/decade), with an        (BoM and CSIRO, 2020)
temperature         increase of 0.16-0.20oC/decade since 1950 in the south-east.
                                                                                      (Perkins-Kirkpatrick et al.,
                    Eight of the ten warmest years on record occurred since 2010.     2016; Alexander and
                                                                                      Arblaster, 2017; Pepler et al.,
Air temperature     More extremely hot days and fewer extremely cold days in          2018; BoM and CSIRO, 2020;
extremes over land  most regions. Weaker warming trends in minimum                    Perkins-Kirkpatrick and
                    temperatures in southeast Australia compared to elsewhere         Lewis, 2020; Trancoso et al.,
                    during 1960-2016. Frost frequency in south-east and south-        2020)
                    west Australia has been relatively unchanged since the 1980s.
                    Very high monthly maximum or minimum temperatures that
                    occurred around 2% of the time in the past (1960­1989) now
                    occur 11-12% of the time (2005­2019). Multi-day heatwave
                    events have increased in frequency and duration across many
                    regions since 1950. In 2019, the national average maximum
                    temperature exceeded the 99th percentile on 43 days (more
                    than triple the number in any of the years prior to 2000) and
                    exceeded 39°C on 33 days (more than the number observed
                    from 1960 to 2018 combined).

Sea temperature     Intense marine heatwave in 2011 near Western Australia (peak      (BoM and CSIRO, 2018;
extremes            intensity 4°C, duration 100 days) - likelihood of an event of     BoM, 2020a; Laufkötter et al.,
                    this duration estimated to be about 5 times higher than under     2020; Oliver et al., 2021)
Rainfall            pre-industrial conditions. Marine heatwave over northern
Rainfall extremes   Australia in 2016 (peak intensity 1.5°C, duration 200 days).
                    Marine heatwave in the Tasman Sea and around southeast
                    mainland Australia and Tasmania from September 2015 to
                    May 2016 (peak intensity 2.5°C, duration 250 days) -
                    likelihood of an event of this intensity and duration has
                    increased about 50-fold. Marine heatwave in the Tasman Sea
                    from November 2017 to March 2018 (peak intensity 3°C,
                    duration 100 days). Marine heatwave on the Great Barrier Reef
                    in 2020 (peak intensity 1.2°C, duration 90 days) (BoM, 2020).

                    Northern Australian rainfall has increased since the 1970s, with  (Delworth and Zeng, 2014;
                    an attributable human influence. April to October rainfall has    Knutson and Zeng, 2018; Dey
                    decreased 16% since the 1970s in south-western Australia          et al., 2019; BoM, 2020c;
                    (partly due to human influence) and 12% from 2000-2019 in         BoM and CSIRO, 2020)
                    south-eastern Australia. Australian-average rainfall was lowest
                    on record in 2019.

                    Hourly extreme rainfall intensities increased by 10­20% in        (Donat et al., 2016; Alexander
                    many locations between 1966­1989 and 1990­2013.                   and Arblaster, 2017; Evans et
                    Daily rainfall associated with thunderstorms increased 13-24%     al., 2017; Guerreiro et al.,
                    from 1979-2016, particularly in northern Australia. Daily         2018; Dey et al., 2019; BoM
                    rainfall intensity increased in the northwest from 1950­2005      and CSIRO, 2020; Bruyère et
                    and in the east from 1911­2014, and decreased in the south-       al., 2020; Dowdy, 2020; Dunn
                    west and Tasmania from 1911­2010.                                 et al., 2020; Gutiérrez et al.,
                                                                                      2021)

Do Not Cite, Quote or Distribute  11-11                                                           Total pages: 151
FINAL DRAFT                       Chapter 11                                    IPCC WGII Sixth Assessment Report
 Drought
                   Major Australian droughts occurred in 1895-1902, 1914-1915,       (Gallant et al., 2013; Delworth
 Windspeed         1937-1945, 1965-1968, 1982-1983, 1997-2009 and 2017-2019.         and Zeng, 2014; Alexander
                   Fewer droughts have occurred across most of northern and          and Arblaster, 2017; Dai and
                   central Australia since the 1970s, more droughts in the south-    Zhao, 2017; Knutson and
                   west since the 1970s, and mixed drought trends in the south-      Zeng, 2018; Dey et al., 2019;
                   east since the late 1990s.                                        Spinoni et al., 2019; BoM,
                                                                                     2020b; Dunn et al., 2020;
                                                                                     Rauniyar and Power, 2020;
                                                                                     BoM, 2021; Seneviratne et al.,
                                                                                     2021)

                   Windspeed decreased 0.067 m/s per decade over land from           (Troccoli et al., 2012; Young
                   1941-2016, with a decrease of 0.062 m/s per decade over land      and Ribal, 2019; Blunden and
                   from 1979­2015, and a decrease of 0.05-0.10 m/s per decade        Arndt, 2020; Azorin-Molina et
                   over land from 1988-2019. Windspeed increased 0.02 m/s per        al., 2021)
                   year across the Southern Ocean from 1985-2018.

Sea-level rise     Relative sea level rise was 3.4 mm/year from 1993-2019,           (Watson, 2020)

                   which includes the influence of internal variability (e.g. ENSO)

                   and anthropogenic greenhouse gases.

Fire               An increase in the number of extreme fire weather days from (Dowdy and Pepler, 2018;

                   July 1950 to June 1985 compared to July 1985 to June 2020, BoM and CSIRO, 2020; van

                   especially in the south and east, partly attributed to climate    Oldenborgh et al., 2021)

                   change. More dangerous conditions for extreme pyro

                   convection events since 1979, particularly in south-eastern

                   Australia. Extreme fire weather in 2019-2020 was at least 30%

                   more likely due to climate change.

Tropical cyclones  Fewer tropical cyclones since 1982, with a 22% reduction in       (Pepler et al., 2015b; Ji et al.,
and other storms   translation speed over Australian land areas from 1949-2016.      2018; Kossin, 2018; BoM and
                   No significant trend in the number of East Coast Lows. From       CSIRO, 2020; Dowdy, 2020;
                   1979­2016, thunderstorms and dry lightning decreased in           ICA, 2021) (Bruyère et al.,
                   spring and summer in northern and central Australia, decreased    2020)
                   in the north in autumn, and increased in the south-east in all
                   seasons. Convective rainfall intensity per thunderstorm
                   increased by about 20% in the north and 10% in the south. An
                   increase in the frequency of large to giant hail events across
                   south-eastern Queensland and north-eastern and eastern New
                   South Wales in the most recent decade. Seven major hail
                   storms over eastern Australia from 2014-2020 and three major
                   floods over eastern Australia from 2019-2021.

Snow               At Spencers Creek (1830 m elevation) in NSW, annual               (Bhend et al., 2012; Fiddes et
                   maximum snow depth decreased 10% and length of snow               al., 2015; Pepler et al., 2015a;
                   season decreased 5% during 2000­2013 relative to 1954­1999.       BoM and CSIRO, 2020)
                   At Rocky Valley Dam (1650 m elevation) in Victoria, annual
                   maximum snow depth decreased 5.7 cm/decade from 1954-
                   2011. At Mt Hotham, Mt Buller and Falls Creek (1638-1760 m
                   elevation), annual maximum snow depth decreased
                   15%/decade from 1988-2013.

Ocean acidification Average pH of surface waters has decreased since the 1880s by (BoM and CSIRO, 2020)
                          about 0.1 (over 30% increase in acidity).

1

2 Table 11.2b: Observed climate change for New Zealand.

Climate variable Observed change                                                     References

Air temperature    Increased by 1.1°C from 1909­2019. Warmest year on record (MfE, 2020a; NIWA, 2020)
                   was 2016, followed by 2018 and 1998 as equal 2nd warmest.

Do Not Cite, Quote or Distribute  11-12                                                          Total pages: 151
FINAL DRAFT                       Chapter 11                  IPCC WGII Sixth Assessment Report

                   Six years between 2013 and 2020 were among New Zealand's
                   warmest on record.

Sea surface        Increased by 0.2°C/decade from 1981­2018.                           (MfE, 2020a)
temperature
Air temperature    Number of frost days (below 0 degrees Celsius) decreased at         (Harrington, 2020; MfE,
extremes           12 of 30 sites, the number of warm days (over 25°C) increased       2020a)
                   at 19 of 30 sites, and the number of heatwave days increased at
Sea temperature    18 of 30 sites during 1972­2019. Increase in the frequency of
extremes           hot February days exceeding the 90th percentile between 1980­
Rainfall           1989 and 2010­2019, with some regions showing more than a
                   five-fold increase.
Rainfall extremes
                   The eastern Tasman Sea experienced a marine heatwave in             (NIWA, 2019; Salinger et al.,
Drought            2017/18 lasting 138 days with a maximum intensity of 4.1°C,         2019b; Salinger et al., 2020;
                   and another marine heatwave in 2018/19 lasting 137 days with        Oliver et al., 2021)
Windspeed          a maximum intensity of 2.8°C.

Sea-level rise     From 1960­2019, almost half of the 30 sites had an increase in      (MfE, 2020a)
                   annual rainfall (mostly in the south) and 10 sites (mostly in the
                   north) had a decrease, but few of the trends are statistically
                   significant. Rainfall increased by 2.8% per decade in
                   Whanganui, 2.1% per decade in Milford Sound and 1.3% per
                   decade in Hokitika. Rainfall decreased by 4.3% per decade in
                   Whangarei and 3.2% per decade in Tauranga.

                   The number of days with extreme rainfall increased at 14 of 30      (MfE, 2020a)
                   sites and decreased at 11 sites from 1960­2019. Most sites
                   with increasing annual rainfall had more extreme rainfall and
                   most sites with decreasing annual rainfall had less extreme
                   rainfall.

                   Drought frequency increased at 13 of 30 sites from 1972­2019        (MfE, 2020a)
                   and decreased at 9 sites. Drought intensity increased at 14 sites,
                   11 of which are in the north, and decreased at 9 sites, 7 of
                   which are in the south.

                   Since 1970, the wind belt has often been shifted to the south of    (MfE, 2020a)
                   New Zealand, bringing an overall decrease in wind-speed over
                   the country. For 1980­2019, the annual maximum wind gust
                   decreased at 11 of the 14 sites that had enough data to calculate
                   a trend, and increased at 2 of the 14 sites

                   Increased 1.8 mm/year from 1900­2018 and 2.4 mm/year from (Bell and Hannah, 2019)
                   1961­2018, mostly due to climate change.

Fire               Six of 28 sites (Napier, Lake Tekapo, Queenstown, Gisborne,         (Pearce, 2018; MfE, 2020a)
                   Masterton, and Gore) had an increase in days with very high or
Tropical cyclones  extreme fire danger from 1997­2019 and 6 sites (Blenheim,
and other storms   Christchurch, Nelson, Tara Hills, Timaru, and Wellington) had
                   a decrease. An increase in fire impacts from 1988­2018
                   included homes lost, damaged, threatened and evacuated.

                   No significant change in storminess. Three major floods and         (MfE, 2020a; ICNZ, 2021)
                   two major hail-storms during 2019-2021.

Do Not Cite, Quote or Distribute  11-13                                                              Total pages: 151
FINAL DRAFT                       Chapter 11                                IPCC WGII Sixth Assessment Report

Snow and ice  From 1978-2019, the snowline rose 3.7 m/year. From 1977 to       (Baumann et al.; Salinger et
              2018, glacier ice volume decreased from 26.6 km3 to 17.9 km3     al., 2019a; Chinn and Chinn,
              (a loss of 33%). From 1978-2016, the area of 14 glaciers in the  2020; MfE, 2020a; Salinger et
              Southern Alps declined 21%. The end-of-summer snowline           al., 2021) (Vargo et al., 2020)
              elevation for 50 glaciers rose 300m from 1949-2019. In the
              Southern Alps, extreme glacier mass loss was at least six times
              more likely in 2011, and ten times more likely in 2018, due to
              climate change.

Ocean acidification Sub-Antarctic ocean off the Otago coast became 7% more     (MfE, 2020a)
                          acidic from 1998­2017.

 1

 2

 3 11.2.2 Projected Climate Change

 4

 5 There are three main sources of uncertainty in climate projections: emission scenarios, regional climate
 6 responses, and internal climate variability (CSIRO and BOM, 2015). Emission scenario uncertainty is
 7 captured in four Representative Concentration Pathways (RCPs) for greenhouse gases and aerosols. RCP2.6
 8 represents low emissions, RCP4.5 medium emissions and RCP8.5 high emissions. Regional climate response
 9 uncertainty and internal climate variability uncertainty are captured in climate model simulations driven by
10 the RCPs.

11

12 Further climate change is inevitable, with the rate and magnitude largely dependent on the emission pathway
13 (IPCC, 2021) (very high confidence). Preliminary projections based on CMIP6 models are described in the
14 IPCC Working Group I Atlas. For Australia, the CMIP6 projections broadly agree with CMIP5 projections
15 except for a group of CMIP6 models with greater warming and a narrower range of summer rainfall change
16 in the north and winter rainfall change in the south (Grose et al., 2020). For New Zealand, the CMIP6
17 projections are similar to CMIP5, but the CMIP6 models indicate greater warming, a smaller increase in
18 summer precipitation and a larger increase in winter precipitation (Gutiérrez et al., 2021).

19

20 Dynamical and/or statistical downscaling offers the prospect of improved representation of regional climate
21 features and extreme weather events (IPCC 2021: Working Group I Chapter 10), but the added value of
22 downscaling is complex to evaluate (Ekström et al., 2015; Rummukainen, 2015; Virgilio et al., 2020).
23 Downscaled simulations are available for New Zealand (MfE, 2018) and various Australian regions (Evans
24 et al., 2020) (IPCC 2021: Working Group I Atlas). Further downscaling was recommended by the Royal
25 Commission into National Natural Disaster Arrangements (CoA, 2020e). Projections for rainfall,
26 thunderstorms, hail, lightning and tornadoes have large uncertainties (Walsh et al., 2016; MfE, 2018).

27

28 Future changes in climate variability are affected by the El Niño Southern Oscillation (ENSO), Southern
29 Annular Mode (SAM), Indian Ocean Dipole (IOD) and Interdecadal Pacific Oscillation (IPO). An increase
30 in strong El Niño and La Niña events is projected (Cai, 2015), along with more extreme positive phases of
31 the IOD (Cai et al., 2018) and a positive trend in SAM (Lim et al., 2016), but potential changes in the IPO
32 are unknown (NESP ESCC, 2020). There is uncertainty about regional climate responses to projected
33 changes in ENSO (King et al., 2015; Perry et al., 2020; Virgilio et al., 2020).

34

35 Australian climate projections are quantified with references in Table 11.3a. Further warming is projected,
36 with more hot days, fewer cold days, reduced snow cover, ongoing sea-level rise and ocean acidification
37 (very high confidence). Winter and spring rainfall and soil moisture are projected to decrease with more
38 droughts in southern Australia, increased extreme rainfall intensity, higher evaporation rates, decreased wind
39 over southern mainland Australia, increased wind over Tasmania, and more extreme fire weather in southern
40 and eastern Australia (high confidence). Increased winter rainfall is projected over Tasmania, with decreased
41 rainfall in south-western Victoria in autumn and in western Tasmania in summer, fewer tropical cyclones
42 with a greater proportion of severe cyclones and decreased soil moisture in the north (medium confidence).
43 Hailstorm frequency may increase (low confidence).

44

45 New Zealand climate projections are quantified with references in Table 11.3b. Further warming is
46 projected, with more hot days, fewer cold days, less snow and glacial ice, ongoing sea-level rise and ocean

Do Not Cite, Quote or Distribute  11-14                                                      Total pages: 151
FINAL DRAFT                       Chapter 11                         IPCC WGII Sixth Assessment Report

 1 acidification (very high confidence). Increases in winter and spring rainfall are projected in the west of the
 2 North and South Islands, with drier conditions in the east and north, caused by stronger westerly winds
 3 (medium confidence). In summer, wetter conditions are projected in the east of both islands, with drier
 4 conditions in the west and central North Island (medium confidence). Fire weather is projected to increase in
 5 most areas, except for Taranaki-Manawat, West Coast and Southland (medium confidence). Extreme
 6 rainfall is projected to increase over most regions, with increased extreme wind-speeds in eastern regions,
 7 especially in Marlborough and Canterbury, and reduced relative humidity almost everywhere, except for the
 8 West Coast in winter (medium confidence). Drought frequency may increase in the north (medium
 9 confidence).

10

11 Table 11.3a: Projected climate change for Australia. Projections are given for different Representative Concentration
12 Pathways (RCP2.6 is low, RCP4.5 is medium, RCP8.5 is high) and years (e.g. 20­year period centered on 2090).
13 Uncertainty ranges are generally 10­90th percentile, and median projections are given in square brackets where
14 possible. The four Australian regions are shown in Chapter 2 of (CSIRO and BOM, 2015). Preliminary projections
15 based on CMIP6 models are included for some climate variables from the IPCC (2021) Working Group 1 report.

Climate variable Projected change (year, RCP) relative to 1986-2005                    References

Air temperature    Annual mean temperature                                             (NESP ESCC, 2020;
                                                                                       IPCC, 2021)
                   · +0.5­1.5°C (2050, RCP2.6), +1.5­2.5°C (2050, RCP8.5), +0.5­
                        1.5°C (2090, RCP2.6), +2.5­5.0°C (2090, RCP8.5)

                   · Weaker increase in the south, stronger increase in the centre.
                   · Preliminary CMIP6 projections: +0.6-1.3oC (2050, SSP1-RCP2.6),

                        +1.2-2.0oC (2050, SSP5-RCP8.5), +0.6-1.5oC (2090, SSP1-
                        RCP2.6), +2.8-4.9oC (2090, SSP5-RCP8.5) relative to 1995-2014

Sea surface        · + 0.4­1.0°C (2030, RCP8.5),                                       (CSIRO and BOM,
temperature        · +2­4°C (2090, RCP8.5).                                            2015)

Air temperature    · Annual frequency of days over 35°C may increase 20­70% by         (CSIRO and BOM,
extremes
                   2030 (RCP4.5), and 25­85% (RCP2.6) to 80­350% (RCP8.5) by 2015; Trancoso et al.,

                   2090                                                                2020)

                   · Heatwaves may be 85% more frequent if global warming increases

                   from 1.5 to 2.0°C, and four times more frequent for a 3°C warming

                   · Annual frequency of frost days may decrease by 10­40% (2030,

                   RCP4.5), 10­40% (2090, RCP2.6) and 50­100% (2090, RCP8.5).

Rainfall           Annual mean rainfall                                                (Liu et al., 2018;
                   · South: ­15 to +2% (2050, RCP2.6), ­14 to +3% (2050, RCP8.5), ­    NESP ESCC, 2020)

                        15 to +3% (2090, RCP2.6), ­26 to +4% (2090, RCP8.5)
                   · East: ­13 to +7% (2050, RCP2.6), ­17 to +8% (2050, RCP8.5), ­

                        19 to +6% (2090, RCP2.6), ­25 to +12% (2090, RCP8.5)
                   · North: ­12 to +5% (2050, RCP2.6), -8 to +11% (2050, RCP8.5), -

                        12 to +3% (2090, RCP2.6), -26 to +23% (2090, RCP8.5)
                   · Rangelands: -18 to +3% (2050, RCP2.6), -15 to +8% (2050,

                        RCP8.5), -21 to +3% (2090, RCP2.6), -32 to +18% (2090,
                        RCP8.5).

Rainfall extremes  Intensity of daily-total rain with 20-year recurrence interval      (NESP ESCC, 2020)
                   · +4 to +10% (2050, RCP2.6),
                   · +8 to +20% (2050, RCP8.5),
                   · +4 to +10% (2090, RCP2.6),
                   · +15 to +35% (2090, RCP8.5).

Drought            Time in drought (Standardized Precipitation Index below -1)         (Kirono et al., 2020)
                   · Southern Australia: 32-46% [39%] (1995), 38-68% [54%] (2050,

                        RCP8.5), 41-81% [60%] (2090, RCP8.5)
                   · Eastern Australia: 25-46% [37%] (1995), 24-67% [47%] (2050,

                        RCP8.5), 19-76% [56%] (2090, RCP8.5)
                   · Northern Australia: 26-44% [34%] (1995), 18-54% [40%] (2050,

                        RCP8.5), 9-81% [39%] (2090, RCP8.5)

Do Not Cite, Quote or Distribute  11-15                                                Total pages: 151
FINAL DRAFT                                  Chapter 11              IPCC WGII Sixth Assessment Report

                    · Australian Rangelands: 29-43% [34%] (1995), 26-58% [42%]
                         (2050, RCP8.5), 23-70% [46%] (2090, RCP8.5).

Windspeed           0-5% decrease over southern mainland Australia and 0-5% increase        (CSIRO and BOM,
Sea-level rise      over Tasmania (2090, RCP8.5)                                            2015)

                    · South (Port Adelaide): 13-29 cm [21 cm] (2050, RCP2.6), 16-33 (McInnes et al., 2015;

                    cm [25 cm] (2050, RCP8.5), 23-55 cm [39 cm] (2090, RCP2.6), Zhang et al., 2017;

                    40-84 cm [61 cm] (2090, RCP8.5)                                         IPCC, 2019b)

                    · East (Newcastle): 14-30 cm [22 cm] (2050, RCP2.6), 19-36 cm [27 (IPCC, 2021)

                    cm] (2050, RCP8.5), 22-54 cm [38 cm] (2090, RCP2.6), 46-88 cm

                    [66 cm] (2090, RCP8.5)

                    · North (Darwin City Council, 2011): 13-28 cm [21 cm] (2050,

                    RCP2.6), 17-33 cm [25 cm] (2050, RCP8.5), 22-55 cm [38 cm]

                    (2090, RCP2.6), 41-85 cm [62 cm] (2090, RCP8.5)

                    · West (Port Hedland): 13-28 cm [20 cm] (2050, RCP2.6), 16-33 cm

                    [24 cm] (2050, RCP8.5), 22-55 cm [38 cm] (2090, RCP2.6), 40-84

                    cm [61 cm] (2090, RCP8.5).

                    These projections have not been updated to include an Antarctic
                    dynamic ice sheet factor which increased global sea level projections
                    for RCP8.5 by ~10 cm. Preliminary CMIP6 projections indicate +40-50
                    cm (2090, SSP1-RCP2.6) and +70-90 cm (2090, SSP5-RCP8.5).

Sea-level extremes  Increase in the allowance for a storm tide event with 1% annual         (McInnes et al., 2015)
                    exceedance probability (100-year return period)
                    · South (Port Adelaide): 21 cm (2050, RCP2.6), 25 cm (2050,

                         RCP8.5), 41 cm (2090, RCP2.6), 66 cm (2090, RCP8.5)
                    · East (Newcastle): 24 cm (2050, RCP2.6), 30 cm (2050, RCP8.5),

                         49 cm (2090, RCP2.6), 86 cm (2090, RCP8.5)
                    · North (Darwin): 21 cm (2050, RCP2.6), 26 cm (2050, RCP8.5), 43

                         cm (2090, RCP2.6), 71 cm (2090, RCP8.5)
                    · West (Port Hedland): 21 cm (2050, RCP2.6), 26 cm (2050,

                         RCP8.5), 43 cm (2090, RCP2.6), 70 cm (2090, RCP8.5).

Fire                · East: annual number of severe fire weather days 0 to +30% (2050, (Clarke and Evans,

Tropical cyclones   RCP2.6), 0 to +60% (2050, RCP8.5), 0 to +30% (2090, RCP2.6), 0 2019; Dowdy et al.,
and other storms
                    to +110% (2090, RCP8.5)                                                 2019, {Clark, 2021

                    · Elsewhere: number of severe fire weather days +5 to +35% (2050, #2658; Virgilio et al.,

                    RCP2.6), +10 to +70% (2050, RCP8.5), +5 to +35% (2090,                  2019; NESP ESCC,

                    RCP2.6) +20 to +130% (2090, RCP8.5).                                    2020)

                    · Eastern region tropical cyclones: -8 to +1% (2050, RCP2.6), -15 to (NESP ESCC, 2020;

                    +2% (2050, RCP8.5), -8 to +1% (2090, RCP2.6), -25 to +5%                Raupach et al., 2021)

                    (2090, RCP8.5)

                    · Western region tropical cyclones: -10 to -2% (2050, RCP2.6), -20

                    to -4% (2050, RCP8.5), -10 to -2% (2090, RCP2.6), -30 to -10%

                    (2090, RCP8.5)

                    · East coast lows: -15 to -5% (2050, RCP2.6), -30 to -10% (2050,

                    RCP8.5), -15 to -5% (2090, RCP2.6), -50 to -20% (2090, RCP8.5).

                    · Hailstorm frequency may increase, but there are large uncertainties.

Do Not Cite, Quote or Distribute                11-16                                       Total pages: 151
FINAL DRAFT                                Chapter 11                IPCC WGII Sixth Assessment Report

Snow and ice     · Maximum snow depth at Falls Creek and Mt Hotham may decline         (Bhend et al., 2012;
                      30­70% (2050, B1) and 45­90% (2050, A1FI) relative to 1990.      Harris et al., 2016; Di
                                                                                       Luca et al., 2018)
                 · Maximum snow depth at Mt Buller and Mt Buffalo may decline
                      40­80% (2050, B1) and 50­100% (2050, A1FI) relative to 1990.

                 · Length of Victorian ski-season may contract 65­90% and mean
                      annual snowfall may decline 60­85% (2070­2099, RCP8.5)
                      relative to 2000­2010.

                 · The snowpack may decrease by about 15% (2030, A2) to 60%
                      (2070, A2).

Ocean acidification pH is projected to drop by about 0.1 (2090, RCP2.6) to 0.3 (2090,  (CSIRO and BOM,
                          RCP8.5).                                                     2015; Hurd et al.,
                                                                                       2018)

1

2 Table 11.3b: Projected climate change for New Zealand. Projections are given for different Representative
3 Concentration Pathways (RCP2.6 is low, RCP4.5 is medium, RCP8.5 is high) and years (e.g. 20­year period centered
4 on 2090). Uncertainty ranges are 5­95th percentile, and median projections are given in square brackets where possible.
5 Preliminary projections (10-90th percentile) based on CMIP6 models are included for some climate variables from the
6 IPCC (2021) Working Group 1 report.

Climate variable Projected change (year, RCP) relative to 1986-2005                    References

Air temperature  Annual mean temperature                                               (MfE, 2018)
                                                                                       (IPCC, 2021)
                 · +0.2­1.3°C [0.7°C] (2040, RCP2.6), +0.5­1.7°C [1.0°C] (2040,
                      RCP8.5), +0.1­1.4°C [0.7°C] (2090, RCP2.6), +2.0­4.6°C [3.0°C]
                      (2090, RCP8.5)

                 · More warming in summer and autumn, less in winter and spring.

                 · More warming in the north than the south.
                 · Preliminary CMIP6 projections: +0.4-1.1oC (2050, SSP1-RCP2.6),

                      +0.9-1.7oC (2050, SSP5-RCP8.5), +0.5-1.5oC (2090, SSP1-
                      RCP2.6), +2.2-4.1oC (2090, SSP5-RCP8.5) relative to 1995-2014

Sea surface      · +1.0°C (2045, RCP8.5),                                              (Law et al., 2018b)
temperature      · +2.5°C (2090, RCP8.5).

Air temperature  · Annual frequency of days over 25°C may increase 20­60% (2040, (MfE, 2018)
extremes              RCP2.6) to 50­100% (2040, RCP8.5), and 20­60% (2090, RCP2.6)
                      to 130­350% (2090, RCP8.5)

                 · Annual frost frequency may decrease 20­60% (2040, RCP2.6) to
                      30­70% (2040, RCP8.5), and 20­60% (2090, RCP2.6) to 70­95%
                      (2090, RCP8.5).

Do Not Cite, Quote or Distribute           11-17                                       Total pages: 151
FINAL DRAFT                                 Chapter 11                        IPCC WGII Sixth Assessment Report

Rainfall            Annual mean rainfall                                                       (Liu et al., 2018;
                    · Waikato, Auckland and Northland: -7 to +7% (2040, RCP2.6), -8 to         MfE, 2018)

                         +5% (2040, RCP8.5), -5 to +11% [+2%] (2090, RCP2.6), -15 to
                         +12% [-2%] (2090, RCP8.5)
                    · Hawke's Bay and Gisborne: -8 to +8% [-1%] (2040, RCP2.6), -12
                         to +7% [-2%] (2040, RCP8.5), -9 to +4% [-2%] (2090, RCP2.6), -
                         15 to +15% [-3%] (2090, RCP8.5)
                    · Taranaki, Manawat and Wellington: -4 to +9% [+1%] (2040,
                         RCP2.6), -6 to +10% [+1%] (2040, RCP8.5), -6 to +15% [+3%]
                         (2090, RCP2.6), -14 to +14% [+2%] (2090, RCP8.5)
                    · Tasman-Nelson and Marlborough: -3 to +5% [+1%] (2040,
                         RCP2.6), -3 to +8% [+1%] (2040, RCP8.5), -4 to +8% [+2%]
                         (2090, RCP2.6), -3 to +15% [+5%] (2090, RCP8.5)
                    · West Coast and Southland: -4 to +12% [+3%] (2040, RCP2.6), -4 to
                         +12% [+4%] (2040, RCP8.5), -2 to +18% [+5%] (2090, RCP2.6), -
                         8 to +23% (2090, RCP8.5)
                    · Canterbury and Otago: -7 to +15% [+3%] (2040, RCP2.6), -7 to
                         +19% [+3%] (2040, RCP8.5), -6 to +18% (2090, RCP2.6), -9 to
                         +28% [+8%] (2090, RCP8.5).

Rainfall extremes   Intensity of daily rain with 20-year recurrence interval                   (MfE, 2018)
                    · +2.8 to 7.2% [5%] (2040, RCP2.6)
                    · +4.2 to 10.4% [7%] (2040, RCP8.5)
                    · +2.8 to 7.2% [5%] (2090, RCP2.6)
                    · +12.6 to 31.5% [2%] (2090, RCP8.5).

Drought             Increase in potential evapotranspiration deficit                           (MfE, 2018)
Windspeed           · Northern and eastern North Island: 100-200 mm (2090, RCP8.5)             (MfE, 2018)
                    · Western North Island: 50-100 mm (2090, RCP8.5)
                    · Eastern South Island: 50-200 mm (2090, RCP8.5)
                    · Western South Island: 0-50 mm (2090, RCP8.5).

                    99th percentile of daily mean wind speed
                    · Northern North Island: 0 to -5% (2090, RCP8.5)
                    · Southern North Island: 0 to +5% (2090, RCP8.5)
                    · South Island: 0 to +10% (2090, RCP8.5).

Sea-level rise      · 23 cm (2050, RCP2.6)                                                     (MfE, 2017a; IPCC,
                    · 28 cm (2050, RCP8.5)                                                     2019b)
                    · 42 cm (2090 RCP2.6)
                    · 67 cm (2090 RCP8.5).

                    These projections have not been updated to include an Antarctic
                    dynamic ice sheet factor which increased global sea-level projections for
                    RCP 8.5 by ~10 cm. Preliminary CMIP6 projections indicate 40-50 cm
                    (2090, SSP1-RCP2.6) and 70-90 cm (2090, SSP5-RCP8.5).

Sea-level extremes  For a rise in sea level of 30 cm, the 1-in-100-year high water levels may  (PCE, 2015)
                    occur about:
                    · Every 4 years at the port of Auckland
                    · Every 2 years at the port of Dunedin
                    · Once a year at the port of Wellington
                    · Once a year at the port of Christchurch.

Do Not Cite, Quote or Distribute            11-18                                              Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report
 Fire
                   · Seasonal Severity Rating (SSR) increases 50-100% in coastal           (Pearce et al., 2011)

                   Marlborough and Otago, 40-50% in Wellington and 30-40% in

                   Taranaki and Whanganui, 0-30% elsewhere (2050, A1B).

                   · Number of days with very high or extreme fire weather increase

                   >100% in coastal Otago, Marlborough and the lower North Island,

                   50-100% in Taup and Rotorua, 20-50% in the rest of the North

                   Island, and little change in the rest of the South Island (2050, A1B).

Tropical cyclones  Poleward shift of mid-latitude cyclones and potential for a small       (MfE, 2018)
and other storms   reduction in frequency.
                                                                                           (Hendrikx et al.,
Snow and ice       · Maximum snow depth on 31 August may decline by 0-10% (2040,           2013; MfE, 2018;
                        A1B) and 26-54% (2090, A1B).                                       Marzeion et al., 2020)
                                                                                           (Anderson et al.
                   · Annual snow days may be reduced by 5-15 days (2040, RCP2.6),          2021)
                        10-25 days (2040, RCP8.5), 5-15 days (2090, RCP2.6) and 15-45
                        days (2090 RCP8.5).

                   · Relative to 2015, New Zealand glaciers are projected to lose 36%,
                        53% and 77% of their mass by the end of the century under
                        RCP2.6, RCP4.5 and RCP8.5, respectively.

                   · Over the period 2006-2099, New Zealand glaciers are projected to
                        lose 50 to 92% of their ice volume for RCP2.6 to RCP8.5.

Ocean              pH is projected to drop by about 0.1 (2090, RCP2.6) to 0.3 (2090        (CSIRO and BOM,
acidification      RCP8.5).                                                                2015; Hurd et al.,
                                                                                           2018; Law et al.,
                                                                                           2018b)

 1

 2

 3 11.3 Observed Impacts, Projected Impacts and Adaptation

 4

 5 This section assesses observed impacts, projected risks, and adaptation for 10 sectors and systems. Boxes
 6 provide more detail on specific issues. Risk is considered in terms of vulnerability, hazards (impact driver),
 7 exposure, reasons for concern, complex and cascading risks (Chapter 1 Figure 1.2).

 8

 9 11.3.1 Terrestrial and Freshwater Ecosystems

10

11 11.3.1.1 Observed Impacts

12

13 Widespread and severe impacts on ecosystems and species are now evident across the region (very high
14 confidence) (Table 11.4). Climate impacts reflect both on-going change and discrete extreme weather events
15 (Harris et al., 2018) and the climatic change signal is emerging despite confounding influences (Hoffmann et
16 al., 2019). Fundamental shifts are observed in the structure and composition of some ecosystems and
17 associated services (Table 11.4). Impacts documented for species include global and local extinctions,
18 severe regional population declines, and phenotypic responses (Table 11.4). In terrestrial and freshwater
19 ecosystems, land use impacts are interacting with climate, resulting in significant changes to ecosystem
20 structure, composition and function (Bergstrom et al., 2021) with some landscapes experiencing catastrophic
21 impacts (Table 11.4). Some of observed changes may be irreversible where projected impacts on ecosystems
22 and species persist (Table 11.5). Of note is the global extinction of an endemic mammal species, the Bramble
23 Cay melomys (Melomys rubicola), from the loss of habitat attributable in part to sea-level rise and storm
24 surges in the Torres Strait (Table 11.4).

25

26 Natural forest and woodland ecosystem processes are experiencing differing impacts and responses
27 depending on the climate zone (high confidence). In Australia, an overall increase in the forest fire danger
28 index, associated with warming and drying trends (Table 11.2a), has been observed particularly for southern
29 and eastern Australia in recent decades (Box 11.1). The 2019-2020 mega wildfires of south eastern Australia
30 burnt between 5.8 - 8.1 million hectares of mainly temperate broadleaf forest and woodland, but with
31 substantial areas of rainforest also impacted, and were unprecedented in their geographic location, spatial

Do Not Cite, Quote or Distribute  11-19                                                    Total pages: 151
FINAL DRAFT                                       Chapter 11     IPCC WGII Sixth Assessment Report

 1 extent, and forest types burnt (Boer et al., 2020; Nolan et al., 2020; Abram et al., 2021; Collins et al., 2021;
 2 Godfree et al., 2021). The human influence on these events is evident (Abram et al., 2021; van Oldenborgh
 3 et al., 2021) (Box 11.1). The fires had significant consequences for wildlife (Hyman et al., 2020; Nolan et al.,
 4 2020; Ward et al., 2020) (Box 11.1) and flow-on impacts for aquatic fauna (Silva et al., 2020). In southern
 5 Australia, deeply rooted native tree species can access soil and ground-water resources during drought,
 6 providing a level of natural resilience (Bell and Nikolaus Callow, 2020; Liu et al., 2020). However, the
 7 Northern Jarrah forests of south western Australia have experienced tree mortality and dieback from long
 8 term precipitation decline and acute heatwave-compounded drought (Wardell-Johnson et al., 2015; Matusick
 9 et al., 2018). While there is limited information on observed impacts for New Zealand, increased mast
10 seeding events in beech forest ecosystems that stimulate invasive population irruptions have been recorded
11 (Schauber et al., 2002; Tompkins et al., 2013).

12

13

14 Table 11.4: Observed impacts on terrestrial and freshwater ecosystems and species in the region where there is
15 documented evidence that these are directly (e.g. a species thermal tolerances are exceeded) or indirectly (e.g. through
16 changed fire regimes) the result of climate change pressures.

Ecosystem             Climate-related Pressure           Impact                         Source

Australia

Forest and woodlands  30-year declining rainfall         Drought-induced canopy         (Matusick et al., 2018;
of southern and                                          dieback across a range of      Hoffmann et al., 2019)
southwestern                                             forest and woodland types
Australia                                                (e.g. northern jarrah)

                      Multiple wildfires in short        Local extirpations and         (Slatyer, 2010; Bowman
                      succession resulting from          replacement of dominant        et al., 2014; Fairman et
                      increased fire risk conditions     canopy tree species and        al., 2016; Harris et al.,
                      including declining winter         replacement by woody           2018; Zylstra, 2018)
                      rainfall and increasing hot days   shrubs due to seeders having
                                                         insufficient time to reach
                                                         reproductive age (Alpine
                                                         Ash) or vegetative
                                                         regeneration capacity is
                                                         exhausted (Snow Gum
                                                         woodlands)

                      Background warming and             Death of fire sensitive trees  (Hoffmann et al., 2019)
                      drying created soil and            species from unprecedented
                      vegetation conditions that are     fire events (Palaeo-endemic
                      conducive to fires being ignited   pencil pine forest growing in
                      by lightning storms in regions     sphagnum, Tasmania, killed
                      that have rarely experienced fire  by lightning-ignited fires in
                      over the last few millennia        2016)

Australia Alps        Severe winter drought; warming     Shifts in dominant vegetation  (Bhend et al., 2012;
Bioregion and         and climate-induced biotic         with a decline in grasses and  Hoffmann et al., 2019)
Tasmanian alpine      interactions                       other graminoids and an
zones                                                    increase in forb and shrub
                                                         cover in Bogong High
                                                         Plains, Victoria, Australia

Do Not Cite, Quote or Distribute                        11-20                                   Total pages: 151
FINAL DRAFT                                         Chapter 11  IPCC WGII Sixth Assessment Report

                      Snow loss, fire, drought and       Changing interactions within    (Hoffmann et al., 2019)
                      temperature changes                and among three key alpine
                                                         taxa related to food supply
                      Retreat of snow line               and vegetation habitat
                                                         resources: The mountain
                                                         pygmy-possum (Burramys
                                                         parvus), the mountain plum
                                                         pine (Podocarpus lawrencei)
                                                         and the bogong moth
                                                         (Agrostis infusia)

                                                         Increased species diversity in (Slatyer, 2010)
                                                         alpine zone

                      Reduced snow cover                 Loss of snow-related habitat    (ACE CRC, 2010; Pepler
                                                         for alpine zone endemic and     et al., 2015a; Thompson,
                                                         obligate species                2016; Mitchell et al.,
                                                                                         2019)

Wet Tropics World     Warming and increasing length      Some vertebrate species have    (Moran et al., 2014;
Heritage Area         of dry season                      already declined in both        Hoffmann et al., 2019)
                                                         distribution area and
Sub-Antarctic                                            population size, both earlier
Macquarie island                                         and more severely than
                                                         originally predicted
Mass mortality of
wildlife species      Reduced summer water               Dieback in the critically       (Bergstrom et al., 2015;
(flying foxes,        availability for 17 consecutive    endangered habitat-forming      Hoffmann et al., 2019)
freshwater fish)      summers, and increases in mean     cushion plant Azorella
                      wind speed, sunshine hours and     macquariensis in the fellfield
                      evapotranspiration over four       and herb field communities
                      decades

                      Extreme heat events; rising        flying foxes - thermal          (AAS, 2019; Ratnayake et
                      water temperatures, temperature    tolerances of species           al., 2019; Vertessy et al.,
                      fluctuations, altered rainfall     exceeded; fish - amplified      2019)
                      regimes including droughts and     extreme temperature
                      reduced in-flows                   fluctuations, increasing
                                                         annual water basin
                                                         temperatures, extreme
                                                         droughts and reduced runoff
                                                         after rainfall

Bramble Cay           Sea-level rise and storm surges    Loss of habitat and global      (Lunney et al., 2014;
                                                         extinction                      Gynther et al., 2016;
melomys (mammal) in Torres Strait                                                        Waller et al., 2017;
                                                                                         CSIRO, 2018)
Melomys rubicola                                                                         (Lunney et al., 2014)

Koala, Phascolarctos  Increasing drought and rising      Population declines and         (Walker et al., 2015)
cinereus              temperatures, compounding          enhanced risk of local
                      impacts of habitat loss, fire and  extinctions
                      increasing human population

Tawny dragon lizard, Desiccation stress driven by        Population decline and
Ctenophorus decresii higher body temperatures and        potential local extinction in
                                                         Flinders Ranges, South
                             declining rainfall          Australia

Do Not Cite, Quote or Distribute                         11-21                           Total pages: 151
FINAL DRAFT                       Chapter 11              IPCC WGII Sixth Assessment Report
 Birds
                    Changing thermal regimes            Changes in body size, mass      (Gardner et al., 2014a;
                    including increasing thermal        and condition and other traits  Gardner et al., 2014b;
                    stress and changes in plant         linked to heat exchange         Campbell-Tennant et al.,
                    productivity are identified causal                                  2015; Gardner et al.,
                                                                                        2018; Hoffmann et al.,
                                                                                        2019)

New Zealand

Forest Birds        Warming                             Increasing invasive predation   (Walker et al., 2019)
Coastal ecosystems                                      pressure on endemic forest
                    More severe storms and rising       birds surviving in cool forest
                    sea levels                          refugia, particularly larger-
                                                        bodied bird species that nest
                                                        in tree cavities and are poor
                                                        dispersers

                                                        Erosion of coastal habitats (Rouse et al., 2017)
                                                        including dunes and cliffs is
                                                        reducing habitat

Beech forest        Increasing mean temperatures        Increased beech mast            (Schauber et al., 2002;
ecosystems          and indirectly through effects of   seeding events that stimulate   Tompkins et al., 2013)
                    events like El Niño­Southern        population irruptions for
                    Oscillation (ENSO)                  invasive rodents and
                                                        mustelids which then prey on
                                                        native species

 1

 2

 3 11.3.1.2 Projected Impacts

 4

 5 In the near-term (2030-2060), climate change is projected to become an increasingly dominant stress on the
 6 region's biodiversity, with some ecosystems experiencing irreversible changes in composition and structure
 7 and some threatened species becoming extinct (high confidence). Climate change will interact with current
 8 ecological conditions, threats and pressures, with cascading ecological impacts, including population
 9 declines, heat-related mortalities, extinctions and disruptions for many species and ecosystems (high
10 confidence) (Table 11.5). These include inadequate allocation of environmental flows for freshwater fish
11 (Vertessy et al., 2019), native forest logging for old-growth forest-dependent fauna (Lindenmayer et al.,
12 2015; Lindenmayer and Taylor, 2020a; Lindenmayer and Taylor, 2020b), and invasive species (Scott et al.,
13 2018). Climate change has synergistic and compounding impacts particularly in bioregions already
14 experiencing ecosystem degradation, threatened endemics, collapse of keystone species, including those of
15 value to Indigenous Peoples, and high extinction rates as a consequence of human activities (Table 11.4)
16 (Gordon, 2009; Australia SoE, 2016; Weeks et al., 2016; Cresswell and Murphy, 2017; Hare et al., 2019;
17 MfE, 2019; Lindenmayer and Taylor, 2020a; Lindenmayer and Taylor, 2020b; Bergstrom et al., 2021). Some
18 native species are projected to have potentially greater geographic range if they can colonise new areas,
19 while other species may be resilient to projected climate change impacts (Bulgarella et al., 2014; K.
20 Lawrence et al., 2017; Conroy et al., 2019; Rizvanovic et al., 2019).

21

22 In southern Australia, some forest ecosystems (alpine ash, snowgum woodland, pencil pine, northern jarrah)
23 are projected to transition to a new state or collapse due to hotter and drier conditions with more fires (Table
24 11.5) (high confidence). In Australia, most native Eucalyptus forest plants have a range of traits that enable
25 them to persist with recurrent fire through recovery buds (sprouters) or regenerate through seeding (Collins,
26 2020), affording them a high level of resilience. For high end projected 2060-2080 fire weather conditions in
27 south east Australia (Clarke and Evans, 2019), stand-killing wildfires could occur at a severity and frequency
28 greater than the regenerative capacity of seeders (Enright et al., 2015; Clarke and Evans, 2019). Most New

Do Not Cite, Quote or Distribute                   11-22                                Total pages: 151
FINAL DRAFT                       Chapter 11                  IPCC WGII Sixth Assessment Report

 1 Zealand native plants are not fire resistant and are projected to be replaced by fire-resistant introduced
 2 species following climate-change related fires (Perry et al., 2014).

 3

 4 A loss of alpine biodiversity in the south-east Australian Alps Bioregion is projected in the near-term due to
 5 less snow on snow patch Feldmark and short alpine herb-fields as well as increased stress on snow-
 6 dependent plant and animal species (high confidence) (Table 11.3, Table 11.5). In Australia, invasive plants
 7 and weeds response rates are expected to be faster than for native species, and climate change could foster
 8 the appearance of a new set of weed species, with many bioregions facing increased impacts from non-native
 9 plants (medium confidence) (Gallagher et al., 2013; Scott et al., 2014; March-Salas and Pertierra, 2020)
10 (Table 11.5), along with declines in some listed weeds (Duursma et al., 2013; Gallagher et al., 2013). In New
11 Zealand, climate change is projected to enable invasive species to expand to higher elevations and
12 southwards (Giejsztowt et al., 2020; MfE, 2020a) (Table 11.5) (medium confidence).

13

14 Projected responses of ecosystem processes are uncertain in part due to complex interactions of climate
15 change with soil respiration, plant nutrient availability (Hasegawa et al., 2015; Orwin et al., 2015; Ochoa-
16 Hueso et al., 2017) and changing fire regimes (Scheiter et al., 2015; Dowdy et al., 2019) (Table 11.5). For
17 aquatic biota, responses will reflect seasonal differences in water temperature (Wallace et al., 2015), and
18 changes in rainfall intensity, productivity and biodiversity (Jardine et al., 2015). Extreme floods may impact
19 negatively on New Zealand river biota by mobilising nutrients, sediments and toxic chemicals, and aiding
20 dispersal of invasive species. These effects are compounded by homogenisation of rivers through
21 channelization (Death et al., 2015).

22

23 Improved coastal modelling, experiments and in situ studies are reducing uncertainties at a local scale about
24 the impact of future sea-level rise on coastal freshwater terrestrial wetlands (medium confidence) (Shoo et
25 al., 2014; Bayliss et al., 2018; Grieger et al., 2019). Low-lying coastal wetlands are susceptible to saltwater
26 intrusion from sea-level rise (Shoo et al., 2014; Kettles and Bell, 2015; Finlayson et al., 2017) with
27 consequences for species dependent on freshwater habitats (Houston et al., 2020). Saline habitat conditions
28 will move inland and new coastal ecosystem states may emerge, including the World Heritage listed
29 Kakadu's freshwater wetland (Bayliss et al., 2018) (Table 11.5). Increasingly, sea-level rise will shrink the
30 intertidal zone, having implications for wading birds which use this zone (Tait and Pearce, 2019) (Box 11.6).
31 The ecology of freshwater wetlands in New Zealand are projected to be impacted by the intersection of
32 warming, drought and heavy rainfall (Pingram et al., 2021) (Table 11.5).

33

34 The impacts on species from projected global warming depend on their physiological and ecological
35 responses for which knowledge is limited (Table 11.5) (Bulgarella et al., 2014; Carter et al., 2018; Green et
36 al., 2021). Knowledge of projected impacts is constrained by uncertainties about the influence of
37 physiological limits, barriers to dispersal, competition, the availability of habitat resources (Worth et al.,
38 2014) and disruptions to ecological interactions (Lakeman-Fraser and Ewers, 2013; Parida et al., 2015;
39 Porfirio et al., 2016). Gaps in ecological modelling of future climate impacts include consideration of long
40 term rainfall and temperature changes (Grimm-Seyfarth et al., 2017; Grimm-Seyfarth et al., 2018), species
41 dispersal rates, evolutionary capacity and phenotypic plasticity and the thresholds at which they are
42 considered adequate to counter the impacts of climate change (Ofori et al., 2017b), as well as indirect effects
43 including sea-level rise and altered fire regimes (Shoo et al., 2014; Cadenhead et al., 2016; He et al., 2016).

44

45

46 Table 11.5: An indicative selection of projected climate-change impacts on terrestrial and freshwater ecosystems and
47 species in Australia and New Zealand respectively.

Ecosystem, species Climate-related pressure Projected Impact  Source

Australia

Do Not Cite, Quote or Distribute  11-23                               Total pages: 151
FINAL DRAFT                                     Chapter 11  IPCC WGII Sixth Assessment Report

Floristic          Increases in temperature     47% of vegetation types have             (Gallagher et al., 2019)
composition of     and reductions in annual     characteristic plant species at risk of
vegetation         precipitation by 2070.       their climatic tolerances being
communities        Many plant species based     exceeded from increasing mean
                   on median projection from    annual temperature by 2070 with
                   five global climate models   only 2% at risk from reductions in
                   (ACCESS1.0, CNRM-            annual precipitation by 2070
                   CM5, HADGEM2-CC,
                   MIROC5, NorESM1-M)
                   centred on the decade 2070
                   under RCP8.5.

Some south east    Reduction in winter          Increase in fire frequency prevents      (Doherty et al., 2017;
Australian         rainfall and rising spring   recruitment of obligate seeder           Zylstra, 2018; Bowman
temperate forests  temperatures resulting in    resulting in changing dominant           et al., 2019; Dowdy et
                   an increase in the           species and vegetation structure         al., 2019; Naccarella et
                   frequency of very high fire  including long lasting or irreversible   al., 2020)
                   weather conditions and       shift in formation from tall wet
                   increased risk of            temperate eucalypt forests dominated     (Wardell-Johnson et al.,
                   catastrophic wildfires;      by obligate seeder trees (e.g. Alpine    2015)
                   based on output from 15      Ash) to open forest or in worst case
                   CMIP5 GCMs using RCP         to shrubland.
                   8.5 for years for 2060­
                   2079 as compared to          Declining rainfall and regolith
                   1990­2009                    drying, more unplanned, intense fires
                                                and declining productivity places
                                                stress on tree growth and
                                                compromises biodiversity in northern
                                                jarrah forest.

                                                Tree line stasis or regression (Snow     (Doherty et al., 2017)
                                                Gum)                                     (Bowman et al., 2019;
                                                                                         Naccarella et al., 2020)

                   Increase in lightning-       Population collapse and severe range     (Doherty et al., 2017)
                   ignited landscape fires      contraction of slow-growing, fire-       (Bowman et al., 2019)
                   along with contracting       sensitive palaeoendemic temperate
                   palaeoendemic refugia due    rainforest species (e.g. Pencil Pine)
                   to warmer and drier
                   climates

                   Rhizosphere responses or     Plant nutrient availability may be       (Hasegawa et al., 2015;
                   accelerated rates of soil    enhanced                                 Ochoa-Hueso et al.,
                   organic matter                                                        2017)
                   decomposition

Alpine ecosystems  Increasing global warming    Loss of alpine vegetation                (Slatyer, 2010;
                   and rising temperatures      communities (snow patch Feldmark         Morrison and Pickering,
                   ongoing reduction in snow    and short alpine herb-fields) and        2013; Pepler et al.,
                   cover and winter rain, and   increased stress on snow-dependent       2015a; Williams et al.,
                   increasing frequency and     plant and animal species; changing       2015; Harris et al.,
                   magnitude of wildfires       suitability for invasive species         2017)

Northern tropical  Rainfall and CO2 effects     Potentially resulting in an in increase (Scheiter et al., 2015)
savannahs                                       in ecosystem carbon storage

Murray-Darling     Drought                      Reduced river flow; mass fish kills      (Grafton et al., 2014;
River Basin        Elevated CO2 levels                                                   AAS, 2019)
                                                Increase plant water use reduces
Unimpaired river                                stream flow                              (Ukkola et al., 2016)
basins

Do Not Cite, Quote or Distribute                11-24                                    Total pages: 151
FINAL DRAFT                                          Chapter 11                  IPCC WGII Sixth Assessment Report

Bearded dragons        Changes in precipitation      P. henrylawsoni and P. microlepidota     (Wilson and Swan,
(lizards), Pogona                                    to gain suitable habitat, P. nullarbor   2017; Silva et al., 2018)
spp.                                                 and P. vitticeps showing the most
                                                     potential loss

Xeric bees             Broad temperate               Climate resilient, only small response (Silva et al., 2018)
                       tolerances, arid climate
                       adapted

Great desert skink     Buffering capacity of         Warming impacts projected to be          (Moore et al., 2018)
Liopholis kintorei     underground                   indirect
                       microclimates, for
                       nocturnal and crepuscular
                       ectotherms

22 narrow range        Projected changes in          Extinction likely within next 20         (Lintermans et al.,
fish species in        rainfall, run-off, air        years                                    2020)
imminent risk of       temperatures and the
extinction             frequency of
                       extreme events (drought,
                       fire, flood) compound risk
                       from other key threats
                       especially invasive species

Freshwater taxa        Changed hydrological          Substantial changes to the               (James et al., 2017)
(freshwater fish,      regimes                       composition of faunal assemblages in
crayfish, turtles and                                Australian rivers well before the end
frogs)                                               of this century, with gains/losses
                                                     balanced for fish but suitable habitat
                                                     area predicted to decrease for many
                                                     crayfish and turtle species and nearly
                                                     all frog species

New Zealand

Modified lowland       Intersection of warming,      Prolonged anoxic conditions in           (Pingram et al., 2021)
wetlands               drought and heavy rainfall    waterways (blackwater events)
                       (ex-tropical cyclones)        leading to mortality of fish (e.g.
Native forests and                                   shortfin eels) and invertebrates, while
lands                                                botulism outbreaks can lead to
                                                     impacts on waterfowl

                       Elevated CO2 levels,          Short-term beneficial effects on         (Ausseil et al., 2019b)
                       warming, increased            carbon storage. Droughts in eastern
                       precipitation.                areas would decrease productivity
                                                     and rates of carbon storage in the
                                                     medium term

                       Increased fire intensity and  Much of the native vegetation has no     (Perry et al., 2014)
                       frequency in hot and dry      fire adaptations causing vulnerability
                       parts of New Zealand          to local extinction due to `interval
                                                     squeeze'

Freshwater rivers      Rainfall variation            Cascading effects of warming,            (Macinnis-Ng et al.,
                                                     drought, floods, and algal blooms        2021)
Three species of                                     compounded by water abstraction
naturalized woody
weeds                  Warming and increased         Increased geographic range               (Sheppard and Stanley,
                       CO2 levels                                                             2014)

Do Not Cite, Quote or Distribute                     11-25                                    Total pages: 151
FINAL DRAFT                                       Chapter 11                  IPCC WGII Sixth Assessment Report

Kauri tree, Agathis  Lower than average           Increased litter fall                  (Macinnis-Ng and
australis            rainfall stimulates a                                               Schwendenmann, 2015)
                     drought-deciduous
                     response in this evergreen
                     species

Windmill palm        Warming                      Increased geographic range             (Aguilar et al., 2017)

New Zealand          Warming                      Enhanced respiration                   (Graham et al., 2014)

tussock grasslands

Invasive species Warming                          Increased invasive species abundance (Tompkins et al., 2013;

                                                  & increased predation on native        Macinnis-Ng et al.,

                                                  species                                2021)

                     Warming                      Expanded ranges of invasive species (Sheppard and Stanley,

                                                  in higher/cooler areas                 2014; Walker et al.,

                                                                                         2019)

                     Warming                      Change in flowering phenology and (Giejsztowt et al., 2020)
                                                  pollination competition

                     Warming                      Increase in invasive plants, insects,  (Macinnis-Ng et al.,
                                                  and pathogens from                     2021)
                                                  subtropical/tropical climates

Tuatara (reptile),   Warming                      Temperature-dependent sex              (Grayson et al., 2014)
Sphenodon                                         determination with more males hatch
punctatus                                         threatening small isolated
                                                  populations

                     Warming                      Increased geographic range             (Carter et al., 2018)

Cattle tick          Warming                      Increased geographic range and risk (K. Lawrence et al.,

                                                  of tick-spread anaemia in cattle       2017)

Brown mudfish, Drought                            Reduced flow regimes associated        (White et al., 2016b;
Neochanna apoda                                   with drought interact with reduced     White et al., 2017)
                                                  habitat due to land use change,
                                                  leading to population declines and
                                                  potential local extinction

Suter's skink        Warming                      Increased suitable range but unclear   (Stenhouse et al., 2018)
                                                  if dispersal is possible because
(lizard) Oligosoma                                habitats are isolated

suteri

Threatened endemic   Fluctuations in total        Heavy rainfall can flood nests and     (Correia et al., 2015)
passerine bird,      precipitation, particularly  kill fledglings while droughts can
Notiomystis cincta   increased and more           cause population-wide reproductive
                     variable rainfall            failure

Feral cats           Warming                      Increased geographic range             (Aguilar et al., 2015b)

 1

 2

 3 11.3.1.3 Adaptation

 4

 5 Managing climate change risks to ecosystems is primarily based on reducing the impact of other
 6 anthropogenic pressures, including invasive species, and facilitating natural adaptation (high confidence).
 7 This approach is most feasible within protected areas on public, private and Indigenous land and sea (Bellard
 8 et al., 2014; Liu et al., 2020) but is also applicable elsewhere (Barnes et al., 2015). Effective strategies
 9 promote ecosystem resilience through changing unsustainable land uses and management practices,
10 increasing habitat connectivity, controlling introduced species, restoring habitats, implementing appropriate

Do Not Cite, Quote or Distribute                  11-26                                         Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 fire management, integrated risk assessment and adaptation planning (B. Frame et al., 2018; Lindenmayer et
 2 al., 2020; Macinnis-Ng et al., 2021). Complementary approaches include ex situ seed banks (Morrison and
 3 Pickering, 2013; Christie et al., 2020).

 4

 5 Best practice conservation adaptation planning is informed by data on key habitats, including refugia, and
 6 restoration that facilitates species movements and employs adaptive pathways (very high confidence) (Guerin
 7 and Lowe, 2013; Reside et al., 2014; Shoo et al., 2014; Keppel et al., 2015; Andrew and Warrener, 2017;
 8 Baumgartner et al., 2018; Harris et al., 2018; Jacobs et al., 2018a; Das et al., 2019; Walker et al., 2019;
 9 Molloy et al., 2020). Landscape planning (Bond et al., 2014; McCormack, 2018) helps reduce habitat loss
10 and facilitates species dispersal and gene flow (McLean et al., 2014; Shoo et al., 2014; Lowe et al., 2015;
11 Harris et al., 2018; McCormack, 2018) and allows for new ecological opportunities (Norman and Christidis,
12 2016). Coastal squeeze is a threat to freshwater wetlands and requires planning for the potential inland shift
13 (Grieger et al., 2019). Adaptations that maintain critical volumes and periodicity of environmental flows will
14 help protect freshwater biodiversity (Yen et al., 2013; Barnett et al., 2015; Wang et al., 2018b) (Box 11.3).

15

16 Adaptation planning for ecosystems and species requires monitoring and evaluation to identify trigger points
17 and thresholds for new actions to be implemented (high confidence) (Tanner-McAllister et al., 2017;
18 Williams et al., 2020). Best planning practice includes keeping options open (Barnett et al., 2015; Dunlop et
19 al., 2016; Finlayson et al., 2017) and updating management plans in light of new information. New insights
20 are emerging into how species' natural adaptive capacities can inform adaptation planning (Llewelyn et al.,
21 2016; Steane et al., 2017; Hoeppner and Hughes, 2019). Physiological limits to adaptation in some species
22 are being identified (Barnett et al., 2015; Sorensen et al., 2016) and where natural responses are not feasible,
23 human-assisted translocations may be warranted (Becker et al., 2013; Chauvenet et al., 2013; Innes et al.,
24 2019) for some species (Ofori et al., 2017a; Ofori et al., 2017b). Legal reform may be needed to better enable
25 climate adaptation for biodiversity conservation that recognises species' natural adjustments to their
26 distributions, and the difficulties in predicting the consequences for ecological interactions and ecosystem
27 services (McCormack, 2018; McDonald et al., 2019).

28

29 Adaptation research priorities include understanding of the interactions and cumulative impacts of existing
30 stressors and climate change, and the implications for managing ecosystems and natural resources (Williams
31 et al., 2020). For Australia, research on implementation strategies for conservation and managing threats,
32 stress and natural assets is a priority (Williams et al., 2020). For New Zealand, understanding how terrestrial
33 ecosystems and species respond to climate change is a priority and where existing stressors are affecting
34 freshwater quantity and quality, in-situ monitoring to detect and evaluate projections of climate change
35 impacts on biodiversity, and a national data repository are lacking (MfE, 2020a). The projected increase in
36 invasive species indicates the importance of a step up in pest management effort to ensure native species
37 persistence as invasive species spread from climate change (Firn et al., 2015). There remains a gap between
38 the knowledge generated, potential adaptation strategies, and their incorporation into conservation
39 instruments (medium confidence) (Graham et al., 2019; Hoeppner and Hughes, 2019), though there is
40 increasing recognition of the need to improve governance and management structures for their
41 implementation (Christie et al., 2020).

42

43

44 [START BOX 11.1 HERE]

45

46 Box 11.1: Escalating Impacts and Risks of Wildfire

47

48 Fire activity depends on weather, ignition sources, land management practices, and fuel flammability,
49 availability and continuity (Bradstock et al., 2014). Increased fire activity in southeast Australia associated
50 with climate change has been observed since 1950 (Abram et al., 2021) but trends vary regionally (Bradstock
51 et al., 2014) (medium confidence). In New Zealand, there has been an increased frequency of major wildfires
52 in plantations (FENZ, 2018) and at the rural-urban interface (Pearce, 2018) (medium confidence). In northern
53 Australia, increased wet season rainfall (Gallego et al., 2017) has increased dry season fuel loads (Harris et
54 al., 2008).

55

56 In Australia, the frequency and severity of dangerous fire weather conditions is increasing, with partial
57 attribution to climate change (very high confidence) (Dowdy and Pepler, 2018; Abram et al., 2021) (11.2.1,

Do Not Cite, Quote or Distribute  11-27       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Figure Box 11.1.1), especially in southern and eastern Australia during spring and summer (Harris and
 2 Lucas, 2019). Although Australia's eucalypt forests and woodlands are fire adapted (Collins, 2020),
 3 increasing intensity and frequency of fires may exceed their resilience due to shorter intervals between high-
 4 severity fires (Bowman et al., 2014; Etchells et al., 2020; Lindenmayer and Taylor, 2020a). Recent fires have
 5 severely impacted eastern rainforests, including significant Gondwana refugia (Abram et al., 2021). In New
 6 Zealand, the trends in very high and extreme fire weather (1997­2019) have not yet been attributed to
 7 climate change (MfE, 2020a).

 8

 9 Fire weather is projected to increase in frequency, severity and duration for southern and eastern Australia
10 (high confidence) and most of New Zealand (medium confidence) (11.2.2), with projected increases in pyro-
11 convection risk for parts of southern Australia (Dowdy et al., 2019) and increased dry-lightning and fire
12 ignition for southeast Australia (Mariani et al., 2019; Dowdy, 2020). Increased fire risk in spring may reduce
13 opportunities for prescribed fuel-reduction burning in some regions (Harris and Lucas, 2019; Di Virgilio et
14 al., 2020). Fuel dryness is a key constraint on wildfire occurrence (Ruthrof et al., 2016). Vegetation change
15 will affect fuel load and fire risk in different areas in complex ways (Watt et al., 2019; Alexandra and Max
16 Finlayson, 2020; Clarke et al., 2020; Sanderson and Fisher, 2020).

17

18 Direct effects of wildfire include death and injury to people and animals, and damage to ecosystems,
19 property, agriculture, water supplies and other infrastructure (Brodison, 2013; Pearce, 2018; de Jesus et al.,
20 2020; Johnston et al., 2020; Maybery et al., 2020). Indirect effects include electricity and communication
21 blackouts leading to cascading impacts on services, infrastructure and communities (Bowman, 2012;
22 Schavemaker and van der Sluis, 2017).

23

24 For New Zealand, there has been recent increased frequency and magnitude of property losses due to
25 wildfire (Pearce, 2018). The 1660ha Port Hills fire in 2017 resulted in the greatest house losses (9) in almost
26 100 years (Langer et al., 2018), but the subsequent 5540ha Lake Ohau fire destroyed 53 houses in 2020
27 (Waitaki District Council, 2020).

28

29 In Australia, between 1987 and 2016, there were 218 deaths, 1,000 injuries, 2,600 people left homeless and
30 69,000 people affected by wildfire (Deloitte, 2017b). Wildfires cost about $1.1 billion per year on average
31 (11.5.2).

32

33 The Australian wildfires of 2019­2020 resulted in 33 deaths, over 3,000 houses destroyed, $2.3 billion in
34 insured losses, and $3.6 billion in losses for tourism, hospitality, agriculture and forestry (CoA, 2020e;
35 Filkov et al., 2020) (Figure Box 11.1.2). Smoke caused a further 429 deaths and 3230 hospitalizations as a
36 result of respiratory distress and illness, with health costs totalling $1.95 billion (Johnston et al., 2020).
37 These fires burnt about 5.8 to 8.1 million hectares of forest in eastern Australia (Ward et al., 2020; Godfree
38 et al., 2021) resulting in the loss or displacement of nearly 3 billion vertebrate animals (CoA, 2020e; Wintle
39 et al., 2020). 114 listed threatened species lost at least 50% of their habitat, and 49 lost 80% (Wintle et al.,
40 2020) among other severe ecological impacts (Hyman et al., 2020). Smoke carried over 4,000 km to New
41 Zealand where it increased snow/glacier melt through darkening surfaces and produced detectable odour (Pu
42 et al. 2021)(Filkov et al., 2020). The fire season of 2019­20 was at least 30% more likely than a century ago
43 due to the influence of climate change (van Oldenborgh et al., 2021). Following the fires, a Royal
44 Commission into National Natural Disaster Arrangements made 80 recommendations, most of which were
45 accepted by government, including establishing a disaster advisory body and a resilience and recovery
46 agency (11.5.2.3) (CoA, 2020e).

47

48 In the face of climate change and the increased cost of fire damage and suppression, there has been
49 considerable investment in fire risk reduction (Table Box 11.1.1). Recent analysis of 8,800 fires in Australia
50 shows resource constraints in response capacity are a barrier to effectively containing fires (Collins et al.,
51 2018b), compounded by lengthened and more extreme fire seasons.

52

53

Do Not Cite, Quote or Distribute  11-28       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1

2 Figure Box 11.1.1: Change in the annual (July to June) number of days that the Forest Fire Danger Index (FFDI)
3 exceeds its 90th percentile from July 1985 to June 2020 relative to July 1950 to June 1985 (BoM and CSIRO, 2020;
4 Abram et al., 2021).

5

6

Do Not Cite, Quote or Distribute  11-29       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1

2 Figure Box 11.1.2: Cascading impacts on people, economic activity, built assets, ecosystems and species arising from
3 the Black Summer fires of 2019­2020 in eastern and southern Australia (Boer et al., 2020; CoA, 2020e; CoA, 2020b;
4 CoA, 2020a; CSIRO, 2020; Filkov et al., 2020; Johnston et al., 2020; Ward et al., 2020; Wintle et al., 2020; Abram et
5 al., 2021; Godfree et al., 2021).

6

7

Do Not Cite, Quote or Distribute  11-30       Total pages: 151
FINAL DRAFT                       Chapter 11                   IPCC WGII Sixth Assessment Report

1 Table Box 11.1.1: Examples of adaptation options and enablers to reduce wildfire risk (Hart and Langer, 2011;
2 Mitchell, 2013; Price et al., 2015; Tolhurst and McCarthy, 2016; Deloitte, 2017b; Miller et al., 2017; Steffen et al.,
3 2017; Kornakova and Glavovic, 2018; Newton et al., 2018; Pearce, 2018; CoA, 2020e; McKemey et al., 2020).

         Land management

Prescribed burning to reduce fuel load close to built assets.

Engagement with Australia's Aboriginal and Torres Strait Islander Peoples to utilise and learn from their fire
management knowledge and skills to assist in management of the landscape and greenhouse gas mitigation.

Locating power lines appropriately or underground and decentralizing power supply to reduce ignitions.

Preventative, community-based interventions to reduce ignitions from arson and accidental fires.

Reduced exposure of new assets through statutory spatial planning and land use regulations, building codes and
building design standards.

          Communications

          Clearer communication of existing exposure and vulnerability to enable informed decisions about risk tolerance and
          management. This should include sites of key biodiversity that are sensitive or susceptible to fire.

          Increased research to understand interactions between fire, fuel, weather, climate and human factors to enhance
          projections of fire occurrence and behaviour.

          Community education and engagement, encouraging house and property maintenance, improving early warning
          systems, more targeted messaging, and increased emergency evacuation planning and sheltering options.

          Infrastructure

          Enhanced training and support for fire-fighters and aerial fire-fighting assets, including sharing of resources
          nationally and internationally to address the increasing overlap of fire seasons which are lengthening across the
          world.

          Nationally consistent response to exceedance of air quality standards.

          Improved governance arrangements to ensure greater accountability and coordination between agencies, sharing of
          data and resources for emergency planning, and greater understanding of risks to critical infrastructure and supply
          chains.

          Development of new systems to augment capability of fire services and technological advances to detect and
          respond to fires.

 4
 5

 6 [END BOX 11.1 HERE]

 7
 8

 9 11.3.2 Coastal and Ocean Ecosystems

10

11 Australia's EEZ covers over 8.1 million km2 of marine territory, including 50,000 km of coastline (Dhanjal-
12 Adams et al., 2016), spanning sub-Antarctic islands in the south to tropical waters in the north. New
13 Zealand's marine territory extends from the sub-tropics to sub-Antarctic waters, encompassing an EEZ of 4
14 million km2, 18,000 km of coastline and 700 smaller islands and islets, in addition to the two main islands
15 (Costello et al., 2010a; MfE, 2016).

16

17 The marine environment is important to the culture, health and well-being of the region's diverse Indigenous
18 Peoples, including those who had sovereign ownership, governance, resource rights, and stewardship over
19 `Sea Country' for many thousands of years before the current sea level stabilised approximately 6000 years
20 ago and before current coastal ecosystems were established (Rist et al., 2019). Marine environments

Do Not Cite, Quote or Distribute  11-31                        Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 contribute A$69 billion per year to Australia's economy (Eadie et al., 2011), and NZ$4 billion per year to
 2 New Zealand's economy (MfE, 2016). They have a high proportion of rare and endemic species (Croxall et
 3 al., 2012) and provide ecosystem services including food production, coastal protection, tourism and carbon
 4 sequestration (Croxall et al., 2012; Kelleway et al., 2017). Half of the species within New Zealand's seas are
 5 endemic (Costello et al., 2010b).

 6

 7 11.3.2.1 Observed Impacts

 8

 9 Climate change is having major impacts on the region's oceans (very high confidence) (Table 11.6) (Law et
10 al., 2016; Sutton and Bowen, 2019). Rising sea surface temperatures have exacerbated marine heatwaves,
11 notably near Western Australia in 2011, the Great Barrier Reef in 2016, 2017 and 2020, and the Tasman Sea
12 in 2015/2016, 2017/2018 and 2018/19 (Table 11.2) (BoM and CSIRO, 2018; AMS, 2019; NIWA, 2019;
13 Salinger et al., 2019b; Sutton and Bowen, 2019; BoM, 2020a; Salinger et al., 2020; Oliver et al., 2021).
14 Temperature anomalies ranged from 1.2-4.0°C and durations ranged from 90-250 days (Table 11.2).

15

16 Ocean carbon storage and acidification has led to decreased surface pH in the region (Table 11.2) including
17 the sub-Antarctic waters off the East Coast of New Zealand's South Island (very high confidence) (Law et
18 al., 2016). The depth of the Aragonite Saturation Horizon has shallowed by 50­100 m over much of New
19 Zealand, which may limit and/or increase the energetic costs of growth of calcifying species (Anderson et
20 al., 2015; Bostock et al., 2015; Mikaloff-Fletcher et al., 2017) (low confidence).

21

22 In the estuaries of south-western Australia, sustained warming and drying trends have caused dramatic
23 declines in freshwater flows of up to 70% since the 1970s, and increased frequency and severity of
24 hypersaline conditions; enhanced water column stratification and hypoxia; and reduced flushing and greater
25 retention of nutrients (Hallett et al., 2017).

26

27 Extensive changes in the life history and distribution of species have been observed in Australia's (very high
28 confidence) (Gervais et al., 2021) and New Zealand's marine systems (medium confidence) (Table 11.6)
29 (Cross-Chapter box MOVING SPECIES in Chapter 5). New occurrences or increased prevalence of disease,
30 toxins and viruses are evident (de Kantzow et al., 2017; Condie et al., 2019), along with heat stress
31 mortalities and changes in community composition (Wernberg et al., 2016; Zarco-Perello et al., 2017;
32 Thomsen et al., 2019). Extreme climatic events in Australia from 2011 to 2017 led to abrupt and extensive
33 mortality of key habitat-forming organisms ­ corals, kelps, seagrasses, and mangroves ­ along over 45% of
34 the continental coastline of Australia (high confidence) (Babcock et al., 2019).

35

36 In 2016 and 2017, the Great Barrier Reef (GBR) experienced consecutive occurrences of the most severe
37 coral bleaching in recorded history (very high confidence) (Box 11.2), with shallow-water reef in the top two
38 thirds of the GBR affected and the severity of bleaching on individual reefs tightly correlated with the level
39 of local heat exposure (Hughes et al., 2018b; Hughes et al., 2019c). Mass mortality of corals from these two
40 unprecedented events resulted in larval recruitment in 2018 declining by 89% compared to historical levels
41 (Hughes et al., 2019b). Southern reefs were also affected by warming, although significantly less than in the
42 north (Kennedy et al., 2018). Coral reefs in Australia are at very high risk of continued negative effects on
43 ecosystem structure and function (Hughes et al., 2019b) (very high confidence), cultural well-being
44 (Goldberg et al., 2016; Lyons et al., 2019) (very high confidence), food provision (Hoegh-Guldberg et al.,
45 2017) (medium confidence), coastal protection (Ferrario et al., 2014) (high confidence) and tourism (Deloitte
46 Access Economics, 2017; Prideaux and Pabel, 2018; GBRMPA, 2019) (high confidence). If bleaching
47 persists, an estimated 10,000 jobs and A$1 billion in revenue would be lost per year from declines in tourism
48 alone (Swann and Campbell, 2016).

49

50 11.3.2.2 Projected Impacts

51

52 Future ocean warming, coupled with periodic extreme heat events, is projected to lead to the continued loss
53 of ecosystem services and ecological functions (high confidence) (Smale et al., 2019), as species further shift
54 their distributions and/or decline in abundance (Day et al., 2018). Compounding climate-driven changes in
55 the distribution of habitat forming species, invasive macroalgae are predicted to exhibit higher growth under
56 all higher pCO2 and lower pH conditions (Roth-Schulze et al., 2018). Corals and mangroves around northern
57 Australia and kelp and seagrass around southern Australia are of critical importance for ecosystem structure

Do Not Cite, Quote or Distribute  11-32       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 and function, fisheries productivity, coastal protection and carbon sequestration; these ecosystem services are
 2 therefore extremely likely2 to decline with continued warming. Equally, many species provide important
 3 ecosystem structure and function in New Zealand's seas including in the deep sea (Tracey and
 4 Hjorvarsdottir, 2019). The future level of sustainable exploitation of fisheries is dependent on how climate
 5 change impacts these ecosystems. Native kelp is projected to further decline in south-eastern New Zealand
 6 with warming seas (Table 11.6). Climate change could affect New Zealand fisheries' productivity
 7 (Cummings et al., 2021), and both ocean warming and acidification may directly affect shellfish culture
 8 (Cunningham et al., 2016; Cummings et al., 2019), and indirectly through changes in phytoplankton
 9 production (Pinkerton, 2017).

10

11 Climate change related temperature and acidification may affect species sex ratios and thus population
12 viability (medium confidence) (Table 11.3) (Law et al., 2016; Tait et al., 2016; Mikaloff-Fletcher et al.,
13 2017). Acidification may alter sex determination (e.g., in the oyster Saccostrea glomerate), resulting in
14 changes in sex ratios (Parker et al., 2018), and may thus affect reproductive success (low confidence).
15 Decreasing river flows (Chiew et al., 2017) are projected to cause periodically open estuaries across south-
16 west Australia to remain closed for longer periods, inhibiting the extent to which marine taxa can access
17 these systems (Hallett et al., 2017) and with warming predicted to constrain activity in some large fish (Scott
18 et al., 2019b). Major knowledge gaps include environmental tolerances of key life stages, sources of
19 recruitment, population linkages, critical ecological (e.g., predator­prey interactions) or phenological
20 relationships, and projected responses to lowered pH (Fleming et al., 2014; Fogarty et al., 2019).

21

22 Black-browed albatrosses breeding on Macquarie Island may be more vulnerable to future climate-driven
23 changes to weather patterns in the Southern Ocean, and potential latitudinal shifts in the sub-Antarctic Front
24 (Cleeland et al., 2019). New Zealand coastal ecosystems face risks from sea-level rise and extreme weather
25 events (MfE, 2020a).

26

27 Nutrient availability and productivity in sub-tropical waters of New Zealand are projected to decline due to
28 increased sea surface temperature and strengthening of the thermocline, but may increase in sub-Antarctic
29 waters, potentially bringing some benefit to fish and other species (low confidence) (Law et al., 2018b). For
30 New Zealand waters as a whole, declines in net primary productivity of 1.2% and 4.5% are projected under
31 RCP4.5 and RCP8.5 respectively by 2100, and declines in primary production of surface waters by an
32 average 6% from the present day under RCP8.5, with sub-tropical waters experiencing the largest decline
33 (Tait et al., 2016).

34

35 The pH of surface waters around New Zealand is projected to decline by 0.33 under RCP 8.5 by 2090 (Tait
36 et al., 2016), and the depth at which carbonate dissolves is projected to be significantly shallower (Mikaloff-
37 Fletcher et al., 2017) affecting the distribution of some species of calcifying cold water corals (Law et al.,
38 2016) (medium confidence). However, model projections suggest that the top of the Chatham Rise may
39 provide temporary refugia for scleractinian stony corals from ocean acidification because the Chatham Rise
40 sits above the aragonite saturation horizon (Anderson et al., 2015; Bostock et al., 2015). For sub-tropical
41 corals, skeletal formation will be vulnerable to the changes in ocean pH with implications for their longer-
42 term growth and resilience (Foster et al., 2015).

43

44 11.3.2.3 Adaptation

45

46 Climate change adaptation opportunities and pathways have been identified across aquaculture, fisheries,
47 conservation and tourism sectors in the region (MacDiarmid et al., 2013; Fleming et al., 2014; MPI, 2015;
48 Jennings et al., 2016; MfE, 2016; Royal Society Te Aprangi, 2017; Ling and Hobday, 2019) and some
49 stakeholders are already autonomously adapting (Pecl et al., 2019). Some fishing and aquaculture industries
50 use seasonal forecasts of environmental conditions, to improve decision making, risk management, and
51 business planning (Hobday et al., 2016) with potential to use 5-yearly forecasts similarly (Champion et al.,
52 2019). Shifts in the distribution, and availability of target species (e.g., oceanic tuna) would impact the

2 In this Report, the following summary terms are used to describe the available evidence: limited, medium, or robust;
and for the degree of agreement: low, medium, or high. A level of confidence is expressed using five qualifiers: very
low, low, medium, high, and very high, and typeset in italics, e.g., medium confidence. For a given evidence and
agreement statement, different confidence levels can be assigned, but increasing levels of evidence and degrees of
agreement are correlated with increasing confidence.

Do Not Cite, Quote or Distribute  11-33       Total pages: 151
    FINAL DRAFT                        Chapter 11                                IPCC WGII Sixth Assessment Report

1 ability of domestic fishing vessels to continue current fishing practices, with potential social and economic
2 adjustment costs (Dell et al., 2015), including disruption to supply chains (Fleming et al., 2014; Plagányi et
3 al., 2014) (Cross-Chapter Box MOVING SPECIES in Chapter 5). Species abundance data are insufficient to
4 enable projections of climate impacts on fishery productivity. However, fishery and aquaculture industries
5 are considering adaptation strategies, such as changing harvests and relocating farms (Pinkerton, 2017).
6 Thus, while climate change is extremely likely to affect the abundance and distribution of marine species
7 around New Zealand, insufficient monitoring means there is limited evidence of ecosystem level change in
8 biodiversity to date, and no quantitative projections of which species may win and lose to climate change
9 (Table 11.6) (Law et al., 2018a; Law et al., 2018b).

10

11

12 Table 11.6: Observed climate-change related changes in the marine ecosystems of Australia and New Zealand.
13 Climate-related impacts have been documented at a range of scales from single species or region-specific studies, to
14 multi-species or community-level changes.

    Type of change           Examples                               Climate-related        Source
                                                                    Pressure

    Australia

    Reduced activity and     Coral trout (Plectropomus leopardus)   Increased              (Johansen et al., 2014; Scott
    increased energetic      one of Australia's most important      temperature            et al., 2017)
    demands                  commercial and recreational tropical   (experimental
                             finfish species                        laboratory study)
                                                                    and ocean warming

    Estuaries warming and    Australian lagoons and rivers          Warming and            (Scanes et al., 2020)
    freshening               warming and decreasing pH at a         reduction in rainfall
                             faster rate than predicted by climate  (leading to reduced
                             models                                 flows and therefore
                                                                    being less
                                                                    frequently open to
                                                                    the sea)

    Changes in life-history  Reduced size of Sydney rock oysters    Limited capacity to (Fitzer et al., 2018)
    traits, behaviour or     (for commercial sale)                  bio mineralize under
    recruitment                                                     acidification
                                                                    conditions

                             Reduced growth in tiger flathead fish Ocean warming           (Morrongiello and Thresher,
                             in equatorward range                                          2015)

                             55% of 335 fish species became         Ocean warming          (Audzijonyte et al., 2020)
                             smaller and 45% became larger as
                             seas warmed around Australia           (over three decades)

                             Rock lobster display reduced           Increased              (Briceño et al., 2020)
                             avoidance of predators at 23           temperature
                             compared to 20                         (experimental
                                                                    laboratory study)

                             Analysis of stress rings in cores of   Heat events            (DeCarlo et al., 2019)
                             corals from the Great Barrier Reef
                             dating back to 1815, found that
                             following bleaching events, the coral
                             was less affected by subsequent
                             marine heatwaves.

                             Mortality and reductions in spawning 2011 marine              (Caputi et al., 2019)
                             stocks of fishery important abalone, heatwave
                             prawns, rock lobsters

    Do Not Cite, Quote or Distribute   11-34                                                       Total pages: 151
FINAL DRAFT                                 Chapter 11                     IPCC WGII Sixth Assessment Report
  New diseases, toxins
                        Recruitment of coral on GBR reduced Warming-driven (Hughes et al., 2019b)

                        to 11% of long-term average             back-to-back global

                                                                bleaching events

                        Green turtle hatchlings from southern   Increased sand       (Jensen et al., 2018)
                        GBR 65-69% female and hatchlings        temperatures
                        from northern GBR 100% female for
                        last two decades

                        First occurrence of the virulent virus  Detected during      (de Kantzow et al., 2017)
                        causing Pacific Oyster Mortality        heatwave
                        Syndrome (POMS), up to 90% of all
                        farmed oysters died in impacted areas

                        Mussels, scallops, oysters, clams,      Warming and          (Hallegraeff and Bolch,
                        abalone and rock lobsters on the east   extension of the     2016)
                        coast of Tasmania found to have high    East Australian
                        levels of Paralytic Shellfish Toxins,   Current
                        originating from a bloom of the
                        harmful Alexandrium tamarense

                        Range expansion of phytoplankton        Warming and          (Hallegraeff et al., 2020)
                        Noctiluca which can be toxic            extension of the
                                                                East Australian
                                                                Current

                        Mortality fish following algal blooms 2013 marine            (Roberts et al., 2019)

                        in South Australia                      heatwave

Changes in species      Range extensions at the poleward        Ocean warming        (Baird et al., 2012; Robinson
distributions           range limit have been detected in:                           et al., 2015; Sunday et al.,
                        Fish, Cephalopods, Crustaceans,                              2015; Ling et al., 2018;
                        Nudibranchs, Urchins, Corals.                                Nimbs and Smith, 2018;
                                                                                     Ramos et al., 2018; Smith et
                        Contractions in range at the            Ocean warming        al., 2019; Caswell et al.,
                        equatorward range edge have been                             2020)
                        detected in: Anemones, Asteroids,
                        Gastropods, Mussels, Algae.                                  (Pitt et al., 2010; Poloczanska
                                                                                     et al., 2011; Smale et al.,
                                                                                     2019)

                        Australia's most southern dominant      Ocean warming        (Tuckett et al., 2017; Ling et
                        reef building coral, Plesiastrea                             al., 2018)
                        versipora in eastern Bass Strait,       Combination of
                        increasing in abundance at the          increased            (Shalders et al., 2018; Teagle
                        poleward edge of the species' range,    temperatures and     et al., 2018)
                        and also in Western Australia           changes in habitat-
                                                                forming algal        (Prahalad and Kirkpatrick,
                        South-west Australia fish               species              2019)
                        assemblages- warm water fish
                        increasing in density at poleward       Wetter and drier
                        edge of distributions and cool-water    climate
                        species decrease in density at
                        equatorward edge of distributions;
                        increase in warm-water habitat
                        forming species leading to reduced
                        habitat for invertebrate assemblages

                        Predicted reduction range of rare
                        Wilsonia humilis herb in Tasmanian
                        saltmarsh but no change in rest of
                        community

Do Not Cite, Quote or Distribute            11-35                                    Total pages: 151
FINAL DRAFT                                Chapter 11                         IPCC WGII Sixth Assessment Report

Changes in abundance    Shift towards a zooplankton              Ocean warming         (Kelly et al., 2016)
                        community dominated by warm-
                        water small copepods in south-east                             (Duke et al., 2017)
                        Australia
                                                                                       (Wahl et al., 2015; Butler et
                        Diebacks of tidal wetland mangroves      2015­2016             al., 2020; Filbee-Dexter and
                                                                 heatwaves             Wernberg, 2020)
                                                                 combined with
                                                                 moisture stress

                        Decline in giant kelp in Tasmania,       Ocean warming &
                        Australia. Less than 10% remaining.      change in East
                        Loss of kelp Australia-wide totalling    Australian Current
                        at least 140,187 ha                      (lower nutrients)

                        Regional loss of seagrass in Shark       High air and water (Strydom et al., 2020)
                        Bay World Heritage Area, Western         temperatures during
                        Australia                                2011 heatwave

                        Increased annual dugong and inshore      Sustained low air     (Meager and Limpus, 2014)
                        dolphin mortality across Queensland      temperature and
                                                                 increased freshwater
                                                                 discharge during
                                                                 high SOI (ENSO)
                                                                 index

                        Predict equatorward decline and          Ocean warming         (Castro et al., 2020)

                        poleward shift of sea urchin in eastern

                        Australia

                        Increasing mortality of Australian fur   Storm surges and      (McLean et al., 2018) (Box
                        seal pups in low-lying colonies          high tides amplified  11.6)
                                                                 by ongoing sea-
                                                                 level rise

Rapid shifts in         Community-wide tropicalization in        Extreme marine        (Vergés et al., 2016;
community               Australian temperate reef                heatwaves led to a    Wernberg et al., 2016)
composition, structure  communities. Temperate species           100-km range
and integrity           replaced by seaweeds, invertebrates,     contraction of
                        corals, and fishes characteristic of     extensive kelp
                        subtropical and tropical waters          forests

                        On-going declines in habitat-forming     Climate-driven shift  (Thomson et al., 2015;
                        seaweeds                                 of tropical           Nowicki et al., 2017; Zarco-
                                                                 herbivores            Perello et al., 2017)
                        Dieback of temperate seagrass in                               (Wernberg et al., 2016)
                        Shark Bay, Australia, subsequently       2011 Marine           (Strydom et al., 2020)
                        replaced by a tropical early             heatwave
                        successional seagrass with seagrass-
                        associated megafauna (sea turtles)
                        declining in health status

                        Increased herbivory by fish on           Change in species (Zarco-Perello et al., 2019)
                        tropicalized reefs of Western            composition due to
                        Australia                                ocean warming

                        No recovery two years after coral        2011 marine           (Bridge et al., 2014)

                        bleaching and macro alga mortality in heatwave

                        western Australia

Do Not Cite, Quote or Distribute           11-36                                       Total pages: 151
FINAL DRAFT                                    Chapter 11                       IPCC WGII Sixth Assessment Report
  New Zealand
                         Mass mortality of particular coral        2016 marine         (Hughes et al., 2018c)
                         species on affected reefs during          heatwave            (Stuart-Smith et al., 2018)
                         heatwaves on the Great Barrier Reef
                         (eastern Australia) led to altered coral
                         reef structure and species
                         composition 8 months later.

                         Community-wide restructuring along 2016 Marine
                         the Great Barrier Reef, one year after heatwave
                         the 2016 mass bleaching event.

Changes in life-history  Alteration of the shell of pua (black     Lowered pH          (Cummings et al., 2019)
                         footed abalone, Haliotis iris) under      (experimental       (Watson et al., 2018;
                         lowered pH (calcite layer thinner,        laboratory study)   McMahon et al., 2020)
                         greater etching of external shell                             (Watson et al., 2018)
                         surface)
                                                                                       (Salinger et al., 2019b)
                         Decline in maximum swimming               Elevated CO2        (Salinger et al., 2019b)

                         performance of kingfish and snapper (experimental

                                                                   laboratory study)

                         Increased mortality and faster growth Increased

                         in juvenile kingfish                      temperature

                         Earlier spawning of snapper in South 2017­2018

                         Island                                    heatwave

Increase in mortality    Heat stress mortality in salmon farms 2017­18 marine

                         off Marlborough, New Zealand,             heatwave

                         where 20 % of the salmon stocks died

Changes in species       Species increasingly caught further       Ocean warming and (Salinger et al., 2019b)
distributions            south, e.g. snapper and kingfish          2017­2018 marine
                                                                   heatwave

                         Non-breeding distribution of New          Climate warming     (Grecian et al., 2016)
                         Zealand nesting seabird (Antarctic
                         Prion) shifting south with long term
                         climate inferred from stable isotopes

                         Less phytoplankton production in          Ocean warming       (Chiswell and Sutton, 2020)
                         Tasman Sea but more on subtropical
                         front

                         Loss of bull kelp (Durvillaea)            2017-18 heatwave    (Salinger et al., 2019b;
                         populations in southern New Zealand       when sea and air    Thomsen et al., 2019;
                         subsequently replaced by the              temperatures        Salinger et al., 2020)
                         introduced kelp Undaria                   exceeded 23 and 30
                                                                   °C respectively

1

2

3 [START BOX 11.2 HERE]

4

5 Box 11.2: The Great Barrier Reef in Crisis

6

7 The Great Barrier Reef ("GBR") is the world's largest coral reef system, comprising 3,863 reefs over an area
8 of 348,700 km2, stretching for 2,300 km. The GBR is a central cornerstone of the beliefs, knowledges, Lores,
9 languages and ways of living for over 70 geographically and culturally diverse Traditional Owner groups

Do Not Cite, Quote or Distribute               11-37                                   Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 spanning the length of the GBR (Dale et al., 2018), and contributes an estimated A$6.4 billion per year (pre
 2 COVID) to the Australian economy, mainly via tourism. As the world's most extensive coral reef ecosystem,
 3 GBR is a globally outstanding and significant entity, with practically the entire ecosystem inscribed as World
 4 Heritage in 1981 (UNESCO, 2021).

 5

 6 The GBR is already severely impacted by climate change, particularly ocean warming, through more
 7 frequent and severe coral bleaching (Hughes et al., 2018b; Hughes et al., 2019c) (very high confidence). The
 8 worst coral bleaching event on record affected over 90% of reefs in 2016 (Hughes et al., 2018b). In the most
 9 northern 700-km-long section of the GBR in which the heat exposure was the most extreme, 50% of the
10 coral cover on reef crests was lost within eight months (Hughes et al., 2018c). Throughout the entire GBR,
11 including the southern third where heat exposure was minimal, the cover of corals declined by 30% between
12 March and November 2016 (Hughes et al., 2018b). In 2017, the central third of the reef was the most
13 severely affected and the back-to-back regional-scale bleaching events has led to an unprecedented shift in
14 the composition of GBR coral assemblages, transforming the northern and middle sections of the reef system
15 (Hughes et al., 2018c) to a highly degraded state (very high confidence). Coral recruitment to the GBR in
16 2018 was reduced to only 11% of the long-term average (Hughes et al., 2019b). A mass bleaching event also
17 occurred in 2020, making it the third event in five years (BoM, 2020a) (Figure Boxes 11.2.1 and 11.2.2).

18

19 Increased heat exposure also affects the abundance and distribution of associated fish, invertebrates and
20 algae (high confidence) (Stuart-Smith et al., 2018). Thus, coral bleaching is an indicator of thermal effects on
21 coral habitat, fauna and flora. Bleaching is expected to continue for the GBR, and Australia's other coral
22 reef systems (virtually certain). Bleaching conditions are projected to occur twice each decade from 2035
23 and annually after 2044 under RCP8.5, and annually after 2051 under RCP4.5 (Heron et al., 2017). Three
24 degrees of global warming would result in over six times the 2016 level of thermal stress (Lough et al.,
25 2018).

26

27 Increases in cyclone intensity projected for this century, and other extreme weather events, will greatly
28 accelerate coral reef degradation (Osborne et al., 2017). Additionally, through interactions between elevated
29 ocean temperature and coastal runoff (nutrient and sediment), extreme weather events may contribute to an
30 increased frequency and/or amplitude of crown of thorn starfish outbreaks (Uthicke et al., 2015), further
31 reducing the spatial distribution of coral.

32

33 Recovery of coral reefs following repeated disturbance events is slow (Hughes et al., 2019b; IPCC, 2019b),
34 and it takes at least a decade after each bleaching event for the very fastest growing corals to recover (high
35 confidence) (Gilmour et al., 2013; Osborne et al., 2017). Estimates of future levels of thermal stress,
36 measured as 'degree heating months' which incorporates both the magnitude and duration of warm season sea
37 surface temperatures (SST) anomalies, suggest that achieving the 1.5°C Paris Agreement target would be
38 insufficient to prevent more frequent mass bleaching events (very high confidence) (Lough et al., 2018),
39 although it may reduce their occurrence (Heron et al., 2017), and occurrences of warming events similar to
40 2016 bleaching could be reduced by 25% (King et al., 2017).

41

42 Tourist motivations for visiting the GBR are changing, with a recent survey finding that two-thirds of
43 tourists were visiting `before it was gone' and a similar number were reporting damage to the reef ­ an
44 example of `last chance tourism' (Piggott-McKellar and McNamara, 2016). The Australian Government is
45 investing A$1.9 billion to support the Great Barrier Reef through science and practical environmental
46 outcomes including reducing other anthropogenic pressures which can suppress natural adaptive capacity
47 (CoA, 2019b; GBRMPA, 2019). However, adaptation efforts on the Great Barrier Reef aimed specifically at
48 climate impacts, for example, coral restoration following marine heatwave impacts (Boström-Einarsson et
49 al., 2020) may slow the impacts of climate change in small discrete regions of the reef, or reduce short-term
50 socio-economic ramifications, but will not prevent widespread bleaching (Condie et al. 2021).

51

52

Do Not Cite, Quote or Distribute  11-38       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1

 2 Figure Box 11.2.1: Top panels show spatial patterns in heat exposure along the Great Barrier Reef in 2016 (left) and
 3 2017 (right), measured from satellites as Degree Heating Weeks (DHW, oC-weeks). Middle panels show the geographic
 4 footprint of recurrent coral bleaching in 2016 (left) and again in 2017 (right), measured by aerial assessments of
 5 individual reefs (adapted from (Hughes et al., 2019c)). Bottom panels display the density of coral recruits (mean per
 6 recruitment panel on each reef), measured over three decades, from 1996 to 2016 (n = 47 reefs, 1,784 panels) (left),
 7 compared to the density of coral recruits in 2018 after the mass mortality of corals in 2016 and 2017 due to the back-to-
 8 back bleaching events (n = 17 reefs, 977 panels) (right). The area of each circle is scaled to the overall recruit density of
 9 spawners and brooders combined. Yellow and blue indicate the proportion of spawners and brooders,
10 respectively (from (Hughes et al., 2019b)).

11

12

Do Not Cite, Quote or Distribute  11-39       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1

 2 Figure Box 11.2.2: Variation in the severity of mass-bleaching episodes recorded on Australia's Great Barrier Reef
 3 over the last four decades (1980­2020). The overall number of reefs surveyed was substantially higher in 1998, 2002,
 4 2016, 2017 and 2020 when aerial surveys were undertaken, whereas the severity of other more localised bleaching
 5 episodes was documented with in-water surveys (adapted from (Pratchett et al., 2021). Extent of bleaching in 2020 was
 6 similar in severity to 2016, but more geographically widespread and included southern reefs.

 7

 8

 9 [END BOX 11.2 HERE]

10

11

12 11.3.3 Freshwater Resources

13

14 Climate change impacts on freshwater resources cascade across people, agriculture, industries and
15 ecosystems (Boxes 11.3 and 11.5). The challenge of satisfying multiple demands with a finite resource is
16 exacerbated by high inter-annual and inter-decadal variability of river flows, particularly in Australia (Chiew
17 and McMahon, 2002; Peel et al., 2004; McKerchar et al., 2010).

18

19 11.3.3.1 Observed Impacts

20

21 Streamflow has generally increased in northern Australia and decreased in southern Australia since the mid-
22 1970s (Zhang et al., 2016) (high confidence). Declining river flows since the mid-1970s in southwest
23 Australia have led to changed water management (WA Government, 2012; WA Government, 2016). The
24 large decline in river flows during the 1997­2009 `Millennium' drought in south-east Australia resulted in
25 low irrigation water allocations, severe water restrictions and major environmental impacts (Potter et al.,
26 2010; Chiew and Prosser, 2011; Leblanc et al., 2012; van Dijk et al., 2013). The drying in southern Australia
27 highlighted the need for hydrological models that adequately account for climate change (Vaze et al., 2010;
28 Chiew et al., 2014; Saft et al., 2016; Fowler et al., 2018). The decline in streamflow was largely due to the
29 decline in cool season rainfall (which has been partly attributed to climate change) (Figure 11.2) (Timbal and
30 Hendon, 2011; Post et al., 2014; Hope et al., 2017; DELWP, 2020) when most of the runoff in southern
31 Australia occurs.

32

33 In New Zealand, precipitation has generally decreased in the north and increased in the southwest (Figure
34 11.2) (Harrington et al., 2014), but it is difficult to ascertain trends in the relatively short streamflow records.
35 Glaciers in New Zealand's southern alps have lost one third of their mass since 1977 (Mackintosh et al.,
36 2017; Salinger et al., 2019b), and glacier mass loss in 2018 was at least ten times more likely to occur with
37 anthropogenic forcing than without (Vargo et al., 2020).

38

39 11.3.3.2 Projected Impacts

40

41 Projections indicate that future runoff in south-east and south-west Australia are likely to decline (median
42 estimate of 20% and 50% respectively, under 2.2°C global average warming) (Figure 11.3) (Chiew et al.,

Do Not Cite, Quote or Distribute  11-40       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1 2017; Zheng et al., 2019). These projections are broadly similar to those reported previously and in AR5
2 (Teng et al., 2012; Reisinger et al., 2014). The range of estimates arises mainly from the uncertainty in
3 projected future precipitation (Table 11.2a).

4

5

 6

 7 Figure 11.3: Projected changes in mean annual runoff for 2046­2075 relative to 1976­2005 for RCP8.5 from
 8 hydrological modelling with future climate projections informed by 42 CMIP5 GCMs. Projections for RCP4.5 are
 9 about three quarters of the above projections. Plots show median projection, and the 10th and 90th percentile range of
10 estimates. The boundaries are based on hydroclimate regions and major drainage basins. Source: (Zheng et al., 2019).

11

12

13 The runoff decline in southern Australia is projected to be further accentuated by higher temperature and
14 potential evapotranspiration (Potter and Chiew, 2011; Chiew et al., 2014), transpiration from tree regrowth
15 following more frequent and severe wildfires (Brookhouse et al., 2013) (Box 11.1), interceptions from farm
16 dams (Fowler et al., 2015), and reduced surface-groundwater connectivity (limiting groundwater discharge
17 to rivers) in long dry spells (Petrone et al., 2010; Hughes et al., 2012; Chiew et al., 2014) (high confidence).
18 In the longer-term, runoff will also be affected by changes in vegetation and surface-atmosphere feedback in
19 a warmer and higher CO2 environment, but the impact is uncertain because of the complex interactions
20 including changes in climate inputs, fire patterns (Box 11.1) and nutrient availability (Raupach et al., 2013;
21 Ukkola et al., 2016; Cheng et al., 2017).

22

23 Climate change is projected to affect groundwater recharge and the relationship between surface waters and
24 aquifers, and through rising sea-levels where groundwater has a tidal signature (PCE, 2015; MfE, 2017a).
25 Groundwater recharge across southern Australia has decreased in recent decades (Fu et al., 2019) and this
26 trend is expected to continue (Barron et al., 2011; Crosbie et al., 2013) (high confidence). Climate change is
27 also projected to impact water quality in rivers and water bodies, particularly through higher temperature and

Do Not Cite, Quote or Distribute  11-41       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 low flows (Jöhnk et al., 2008) (Box 11.5) and increased sediment and nutrient load following wildfires
 2 (Biswas et al., 2021) (Box 11.1) and floods (Box 11.4) (high confidence).

 3

 4 The projected changes in river flows in New Zealand are consistent with the precipitation projections (Table
 5 11.2), with increases in the west and south of the South Island and decreases in the east and north of the
 6 North Island (Figure 11.4). In the South Island, the runoff increase occurs mainly in winter due to increasing
 7 moisture-bearing westerly airflow, with more precipitation falling as rain and snow melting earlier. In the
 8 North Island, the runoff decrease occurs in spring and summer (Caruso et al., 2017; Collins et al., 2018a;
 9 Jobst et al., 2018; D. Collins, 2020).

10

11

12

13 Figure 11.4: Projected percentage change in mean annual runoff for 2086­2099 relative to 1986­2005 from
14 hydrological modelling informed by six CMIP5 GCMs for four RCPs. Maps show median projection from the six
15 modelling runs. White indicates that the change is not statistically significant. Source: (D. Collins, 2020).

16

17

18 11.3.3.3 Adaptation

19

20 In Australia, prolonged droughts and projections of a drier future have accelerated policy and management
21 change in urban and rural water systems. Adaptation initiatives and mechanisms, like significant government
22 investment to enhance the Bureau of Meteorology online water information (Vertessy, 2013; BoM, 2016),
23 funding to improve agriculture water use and irrigation efficiency (Koech and Langat, 2018), enhanced
24 supply through inter-basin transfers and upgrading water infrastructure, and an active water trading market
25 (Wheeler et al., 2013; Kirby et al., 2014; Grafton et al., 2016) are helping to buffer regional systems against
26 droughts, and facilitating some adaptation to climate change (medium confidence). However, these measures
27 could also be maladaptive as they may perpetuate unsustainable water and land uses under ongoing climate
28 change (Boxes 11.3 and 11.5).

29

30 The widespread 2017-2019 drought across eastern Australia (BoM, 2019) has led to the Australian
31 Government establishing a Future Drought Fund (Australian Government, 2019) to enhance drought
32 resilience, and a National Water Grid Authority to develop regional water infrastructure to support
33 agriculture. Nevertheless, the ability to adapt to climate change is compounded by uncertainties in future
34 water projections, complex interactions between science, policy, community values and political voice, and
35 competition between different sectors dependent on water (Boxes 11.3 and 11.5). The impact of declining
36 water resource on agricultural, ecosystems and communities in south-eastern Australia would escalate with
37 ongoing climate change (Hart, 2016; Moyle et al., 2017) (medium confidence), highlighting the importance
38 of more ambitious, anticipatory, participatory and integrated adaptation responses (Bettini et al., 2015; Abel
39 et al., 2016; Marshall and Lobry de Bruyn, 2021).

Do Not Cite, Quote or Distribute  11-42       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1

 2 Altered water regimes resulting from the combined effects of climatic conditions and water policies carry
 3 uneven and far-reaching implications for communities (high confidence). Acting on Indigenous Peoples'
 4 claims to cultural flows (to maintain connection to Country) is increasingly recognised as an important water
 5 management and social justice issue (Taylor et al., 2017; Hartwig et al., 2018; Jackson, 2018; Jackson and
 6 Moggridge, 2019; Moggridge et al., 2019). Compounding stressors such as coal and coal seam gas
 7 developments can also severely impact local communities, water catchments and water-dependent
 8 ecosystems and assets, exacerbating their vulnerability to climate change (Navi et al., 2015; Tan et al., 2015;
 9 Chiew et al., 2018).

10

11 In Australian capital cities and regional centres, water planning has focused on securing new supplies that
12 are resilient to climate change. This includes increasing use of stormwater and sewage recycling and
13 managed aquifer recharge (Bekele et al., 2018; Page et al., 2018; Gonzalez et al., 2020). All major coastal
14 Australian cities have desalination plants. Household scale adaptation like rainwater harvesting, water smart
15 gardens, dual flush toilets, water-efficient showerheads and voluntary residential use targets can help reduce
16 water demand by up to 40% (Shearer, 2011; Rhodes et al., 2012; Moglia et al., 2018). Water utilities across
17 Australia have established climate change adaptation guidelines (WSAA, 2016). Coordinated efforts to
18 reduce demand, design and retrofit infrastructure to reduce flood risk and harvest water, and water sensitive
19 urban design, are evident (WSAA, 2016; Kunapo et al., 2018; Rogers et al., 2020b). Transitioning
20 centralised water systems to a more sustainable basis represents adaptation progress but is complex and faces
21 many barriers and limits (Morgan et al., 2020) (medium confidence). Developing multiple redundant or
22 decentralised systems can enhance community resilience and promote autonomous adaptations that may be
23 more sustainable and cost effective in the longer term (Mankad and Tapsuwan, 2011; WSAA, 2016; Iwanaga
24 et al., 2020).

25

26 In New Zealand, many water supplies are at risk from drought, extreme rainfall events and sea-level rise,
27 exacerbated by underinvestment in existing water infrastructure (in part due to funding constraints), and
28 urban densification (CCATWG, 2017; MfE and StatsNz, 2021) (high confidence). Lessons can be learned
29 from global experience (e.g. Cape Town, South Africa; Chapter 4.3.4). Water quality has diminished, with
30 hotter conditions and drought causing algal blooms, combined with intensification of agricultural land uses
31 in some areas, and heavy rainfall and sea-level rise causing flooding and sedimentation of water sources and
32 health impacts (11.3.6; Box 11.5). Some towns are only partially metered or not metered at all, which
33 exacerbates the adaptation challenge (Hendy et al., 2018; WaterNz, 2018; Paulik et al., 2019b). Unregulated
34 or absent water supplies accentuate risks to vulnerable groups of people (MfE, 2020b). Mori view water as
35 the essence of all life, which makes any impacts on water, of governance and stewardship concern, and
36 increasingly, the subject of legal claims (MfE, 2020a; MfE, 2020b; MfE, 2020c) (11.4.2). Mori
37 understanding of time can also open up new spaces for rethinking freshwater management in a climate
38 change context that does not reinforce or rearticulate multiple environmental injustices (Parsons et al., 2021).

39

40 Water resource adaptation in New Zealand is variable across local government and water authorities but they
41 all actively monitor water availability, demand and quality, and most have drought management plans. The
42 2019/20 drought led to water shortages in the most populated areas of Waikato, Auckland and Northland,
43 resulting in water reduction advisories and five to eight weeks waiting time for water tank refills and water
44 rationing. The Havelock North water supply contamination that arose after an extreme rainfall event (DIA,
45 2017a; DIA, 2017b) was exacerbated by fragmented governance, and led to the Taumata Arawai-Water
46 Services Regulator Act 2020 and the Water Services Bill 2020 to protect source water. The 2017 update to
47 the National Policy Statement for Freshwater Management with guidelines for implementation at the
48 regional level (MfE, 2017b), including consideration of climate change which creates opportunities for
49 adaptation. However, there remain tensions between land, water and people which are exacerbated by
50 climate changes and yet to be addressed (Box 11.5). The first National Adaptation Plan and the Resource
51 Management law reform have potential for helping to resolve these tensions (11.7.1) (CCATWG, 2017;
52 MfE, 2020a).

53

54

55 [START BOX 11.3 HERE]

56

Do Not Cite, Quote or Distribute  11-43       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Box 11.3: Drought, Climate Change, and Water Reform in the Murray-Darling Basin

 2

 3 The Murray-Darling Basin (MDB) is Australia's largest, most economically important and politically
 4 complex river system (Figure Box 11.3.1). The MDB supports agriculture worth A$24 billion/year, 2.6
 5 million people in diverse rural communities, and important environmental assets including 16 Ramsar listed
 6 wetlands (DAWE, 2021). Climate change is projected to substantially reduce water resources in the MDB
 7 (high confidence), with the median projection indicating a 20% decline in average annual runoff under 2.2°C
 8 average global warming (Figure 11.3) (Whetton and Chiew, 2020). This reduction, plus increased demand
 9 for water in hot and dry conditions, would increase the already intense competition for water (high
10 confidence) (CSIRO, 2008; Hart, 2016).

11

12 The economic, environmental and social impacts of the 1997­2009 `Millennium Drought' in the MDB
13 (Chiew and Prosser, 2011; Leblanc et al., 2012; van Dijk et al., 2013), and projections of a drier future under
14 climate change, have accelerated significant water policy reforms, costing more than A$12 billion (Bark et
15 al., 2014; Docker and Robinson, 2014; Hart, 2016). These reforms included the development of a Basin Plan
16 (MDBA, 2011; MDBA, 2012) requiring consistent regional water resource plans (MDBA, 2011; MDBA,
17 2012; MDBA, 2013) and environmental watering strategies (MDBA, 2014) across the MDB. Despite
18 contestation, the reforms have resulted in some substantive achievements, including returning an equivalent
19 of about one fifth of consumptive water to the environment through the purchase of irrigation water
20 entitlements and infrastructure projects (Hart, 2016; Gawne et al., 2020; MDBA, 2020) (medium
21 confidence). However, the overall impacts of these water management initiatives are difficult to measure due
22 to hydroclimatic variability, time lags and environmental, social and institutional complexity (Crase, 2011;
23 Bark et al., 2014; Docker and Robinson, 2014; MDBA, 2020).

24

25 Reform initiatives such as water markets, improving agriculture water use efficiency (Koech and Langat,
26 2018), and increasing environmental water are helping buffer the system against droughts (Moyle et al.,
27 2017) (medium confidence) but they can also be maladaptive by perpetuating unsustainable water and land
28 use under ongoing climate change. While water markets can allow users to adapt and shift water to higher
29 value uses, they can also have adverse impacts unless supported by wider policy goals and planning
30 processes (Wheeler et al., 2013; Kirby et al., 2014; Grafton et al., 2016; Qureshi et al., 2018).

31

32 Adapting MDB management to climate risks is an escalating challenge, with the projected decline in runoff
33 being potentially greater than the water recovered for the environment (Chiew et al., 2017). While the Basin
34 Plan includes mechanisms for climate risks management (Neave et al., 2015), it does not require altering pre-
35 existing rules that distribute the impacts of anticipated reductions in water resource between users (Hart,
36 2016; Capon and Capon, 2017; Alexandra, 2020). The intense drought conditions in 2017-2019 (BoM,
37 2019), the South Australian Royal Commission into the MDB reforms (SA Government, 2019b), and major
38 fish kills in the lower Darling River in the summer of 2018/2019 (AAS, 2019; Vertessy et al., 2019) have
39 increased concerns about the Basin Plan's climate adaptation deficit (medium confidence). The Murray
40 Darling Basin Authority (MDBA) consequently is undertaking an assessment of climate change risks and
41 developing adaptation mechanisms (MDBA, 2019) that can feed into the revisions to the Basin Plan
42 scheduled for 2026. The MDB reforms to date illustrate the difficulties in integrating climate change science
43 and projections into management (Alexandra, 2018; Alexandra, 2020). Anticipatory and participatory
44 governance and adaptive management approaches supported by structural and institutional reforms would
45 support the effectiveness of the reforms (Abel et al., 2016; Alexandra, 2019; Hassenforder and Barone, 2019;
46 Marshall and Lobry de Bruyn, 2021).

47

48

Do Not Cite, Quote or Distribute  11-44       Total pages: 151
      FINAL DRAFT                 Chapter 11  IPCC WGII Sixth Assessment Report

1

 2

 3 Figure Box 11.3.1: (A) The Murray-Darling Basin, and (B) average annual river flows in the Basin under pre-
 4 development conditions (from (CSIRO, 2008) showing that most of the runoff comes from the south-eastern highlands.
 5 The borders show key drainage basins.

 6

 7

 8 [END BOX 11.3 HERE]

 9

10

11 [START BOX 11.4 HERE]

12

13 Box 11.4: Changing Flood Risk

14

15 Pluvial (flash flood from high intensity rainfall) and fluvial (river) flooding are the most costly natural
16 disasters in Australia, averaging A$8.8 billion per year (Deloitte, 2017b). In New Zealand, insured damages
17 for the 12 costliest flood events from 2007-2017 exceeded NZ$472 million of which NZ$140 million has
18 been attributed to anthropogenic climate change (Frame et al., 2020). Extreme rainfall intensity in northern
19 Australia and New Zealand has been increasing, particularly for shorter (sub-daily) duration and more
20 extreme high rainfall (high confidence) (Westra and Sisson, 2011; Griffiths, 2013; Laz et al., 2014; Rosier et
21 al., 2015). Changes are also occurring in spatial and temporal patterns and seasonality (Wasko and Sharma,
22 2015; Zheng et al., 2015; Wasko et al., 2016).

23

24 Extreme rainfall is projected to become more intense (high confidence) but the magnitude of change is
25 uncertain (Evans and McCabe, 2013; Bao et al., 2017) (Table 11.3). The insured damage in New Zealand
26 from more intense extreme rainfall under RCP8.5 is projected to increase 25% by 2080­2100 (Pastor-Paz et
27 al., 2020). In urban areas, extreme rainfall intensity is projected to increase pluvial flood risk (high
28 confidence). In New Zealand, 20,000km2 of land, 675,000 people, and 411,000 buildings with a NZ$135
29 billion replacement value are exposed to 1-in-100 year flood risk (Paulik et al., 2019a).

30

31 In non-urban areas, where the flood response is also dependent on antecedent catchment conditions (Johnson
32 et al., 2016; Sharma et al., 2018), there is no evidence of increasing flood magnitudes in Australia (Ishak et
33 al., 2013; Zhang et al., 2016; Bennett et al., 2018) except for the most extreme events (Sharma et al., 2018;
34 Wasko and Nathan, 2019). Modelling studies project increases in flood magnitudes in northern and eastern
35 Australia, and in western and northern New Zealand (high confidence) (Hirabayashi et al., 2013; Collins et
36 al., 2018a; Do et al., 2020). The change in flood magnitude in southern Australia is uncertain because of the
37 compensating effect of more intense extreme rainfall, versus projected drier antecedent conditions (Johnson
38 et al., 2016; Pedruco et al., 2018; Wasko and Nathan, 2019). Higher rainfall intensity and peak flows also
39 increase erosion, sediment and nutrient loads in waterways (Lough et al., 2015) and exacerbate problems

Do Not Cite, Quote or Distribute  11-45       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 from aging stormwater and wastewater infrastructure (Jollands et al., 2007; WSAA, 2016; Hughes et al.,
 2 2021).

 3

 4 There is some recognition of the need for flood management and planning to adapt to climate change
 5 (COAG, 2011; CCATWG, 2018; CoA, 2020d) (medium confidence). Australian flood estimation guidelines
 6 recommend a 5% increase in design rainfall intensity per degree global average warming (Bates et al., 2015).
 7 In New Zealand, the recommended increase ranges from 5% to more than 10% for shorter duration and
 8 longer return period storms (MfE, 2010; Carey-Smith et al., 2018). Both guidelines also indicate the
 9 potential for higher increases in extreme rainfall intensity.

10

11 Adaptation to reduce flooding and its impacts have included: improved flood forecasting (Vertessy, 2013;
12 BoM, 2016) and risk management (AIDR, 2017); accommodating risk through raising floor levels and
13 sealing external doors (Queensland Government, 2011; Wang et al., 2015), deployment of temporary levee
14 structures; and risk reduction through spatial planning and relocation. Adaptation options in urban areas
15 include improved stormwater management (Hettiarachchi et al., 2019; Matteo et al., 2019), ecosystem-based
16 approaches such as maintaining floodplains, restoring wetlands and retrofitting existing flood control
17 systems to attenuate flows, and water sensitive urban design (WSAA, 2016; Radcliffe et al., 2017;
18 Radhakrishnan et al., 2017; Rogers et al., 2020b).

19

20 Adaptation to changing flood risks is currently mostly reactive and incremental in response to flood and
21 heavy rainfall events (high confidence). For example, the 2010-2011 flooding in eastern Australia resulted in
22 changes to reservoir operations to mitigate floods (QFCI, 2012) and insurance practice to cover flood
23 damages (Phelan, 2011; Phelan et al., 2011; QFCI, 2012; Schuster, 2013). Nevertheless, adaptation planning
24 that is pre-emptive and incorporates uncertainties into flood projections is emerging (Schumacher, 2020)
25 (medium confidence). Examples from New Zealand include the use of Dynamic Adaptive Pathways Planning
26 (Lawrence and Haasnoot, 2017) with Real Options assessment (Infometrics and PSConsulting, 2015) and
27 design of decision signals and triggers to monitor changes before physical and coping thresholds are reached
28 (Stephens et al., 2018). Implementing adaptive flood risk management relies upon an understanding of how
29 such risks change in uncertain and ambiguous ways necessitating adaptive and robust decision making
30 processes. These can enable learning through participatory adaptive pathways approaches (Lawrence and
31 Haasnoot, 2017; Bosomworth and Gaillard, 2019) and through coordination across different levels of
32 government and statutory mandates, adaptation funding, and individual and community adaptations
33 (Glavovic et al., 2010; Boston and Lawrence, 2018; McNicol, 2021).

34

35 [END BOX 11.4 HERE]

36

37

38 11.3.4 Food, Fibre, Ecosystem Products

39

40 The food, fibre and ecosystem products sectors are economically important in the region. Agriculture
41 contributes around 4% of New Zealand GDP and 2% of Australian GDP, and over 50% of New Zealand's
42 and 11% of Australia's exports (NZ Treasury, 2016; Jackson et al., 2020). Forestry contributes 1% of New
43 Zealand GDP and 0.5% Australian GDP (NZ Treasury, 2016; Whittle, 2019). With the processing and
44 indirect effects, the primary sector of New Zealand contributes 25% of GDP (Saunders et al., 2016). The
45 region has the lowest level of agricultural subsidies across the OECD (OECD, 2017), and highly responsive
46 producers to market drivers but limited strategic, longer-term approaches to environmental challenges and
47 adaptation (Wreford et al., 2019). Both countries receive government financial drought assistance (Pomeroy,
48 2015; Downing et al., 2016).

49

50 Impacts resulting from climate change are observed across sectors and the region (high confidence). While
51 more intense changes are observed in Australia, New Zealand is also experiencing impacts, including the
52 economic impacts of drought attributable to climate change (Frame et al. 2020). Overall, modelling
53 indicates that negative impacts will intensify with increased levels of warming in both countries, with
54 declining crop yield and quality, and negative effects on livestock production and forestry. Although benefits
55 are identified, particularly in the short term for New Zealand (MfE, 2020a), an absence of studies that
56 consider the totality of climatic variables, including extremes, moderate the benefits identified from
57 considering only selected variables and systems in isolation.

Do Not Cite, Quote or Distribute  11-46       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Incremental adaptation is occurring (Hochman et al., 2017; Hughes and Lawson, 2017; Hughes and Gooday,
 2 2021). In the longer term, transformative adaptation, including land-use change, will be required (Cradock-
 3 Henry et al., 2020a), both as a result of sectoral adaptations and mitigation (Grundy et al., 2016) (medium
 4 confidence). Specific changes are context specific and challenging to project (Bryan et al., 2016). Future
 5 adaptive capacity may be limited by declining institutional and community capacity resulting from high debt,
 6 unavailability of insurance, increasing regulatory requirements, and funding mechanisms that lock-in
 7 ongoing exposure to climate risk, creating mental health impacts (Rickards et al., 2014; Wiseman and
 8 Bardsley, 2016; McNamara and Buggy, 2017; McNamara et al., 2017; Moyle et al., 2017; Robinson et al.,
 9 2018; Ma et al., 2020; Yazd et al., 2020).

10

11 11.3.4.1 Field Crops and Horticulture

12

13 11.3.4.1.1 Observed impacts
14 Drought, heat and frost in recent decades have shown the vulnerability of Australian field crops and
15 horticulture to climate change (Cai et al., 2014; Howden et al., 2014; CSIRO and BOM, 2015; Lobell et al.,
16 2015; Hughes and Lawson, 2017; King et al., 2017; Webb et al., 2017; Harris et al., 2020) as recognised by
17 policy makers (CoA, 2019a) (high confidence). Drought has caused economic losses attributable to climate
18 change of at least NZ$800M in New Zealand (Frame et al., 2020). Northern Australia's agricultural output
19 losses are on average 19% each year due to drought (Thi Tran et al., 2016). In southern Australia, the
20 frequency of frost has been relatively unchanged since the 1980s (Dittus et al., 2014; Pepler et al., 2018;
21 BoM and CSIRO, 2020). Drier winters have increased the irrigation requirement for wine grapes (Bonada et
22 al., 2020) while smoke from the 2019/20 fires, which occurred early in the season, caused significant taint
23 damage (Jiang et al., 2021). In New Zealand, reduced winter chill has a compounded impact on the kiwifruit
24 industry, resulting in early harvest and increased energy demand for refrigeration and port access problems
25 (Cradock-Henry et al., 2019) (11.5).

26

27 Across all types of agriculture, drought and its physical flow-on effects have caused financial and emotional
28 disruption and stress in farm households and communities (Austin et al., 2018; Bryant and Garnham, 2018;
29 Yazd et al., 2019) (11.3.6). Severe and uncertain climate conditions are statistically associated with increases
30 in farmer suicide (Crnek-Georgeson et al., 2017; Perceval et al., 2019). Rural women often carry extra stress
31 and responsibilities, including increased unpaid and paid work and emotional load (Whittenbury, 2013;
32 Hanigan et al., 2018; Rich et al., 2018).

33

34 11.3.4.1.2 Projected impacts
35 Australian crop yields are projected to decline due to hotter and drier conditions, including intense heat
36 spikes (Anwar et al., 2015; Lobell et al., 2015; Prokopy et al., 2015; Dreccer et al., 2018; Nuttall et al., 2018;
37 Wang et al., 2018a) (high confidence). Interactions of heat and drought could lead to even greater losses than
38 heat alone (Sadras and Dreccer, 2015; Hunt et al., 2018). Australian wheat yields are projected to decline by
39 2050, with a median yield decline of up to 30% in south-west Australia and up to 15% in South Australia,
40 with possible increases and decreases in the east (Taylor et al., 2018, Wang, #1599; Wang et al., 2018a). In
41 temperate fruit, accumulated winter chill for horticulture is projected to further decline (Darbyshire et al.,
42 2016). Winegrape maturity is projected to occur earlier due to warmer temperatures (Webb et al., 2014; van
43 Leeuwen and Darriet, 2016; Jarvis et al., 2018; Ausseil et al., 2019b) (high confidence) leading to potential
44 changes in wine style (Bonada et al., 2015). Rice is susceptible to heat stress and average grain yield losses
45 across rice varieties range from 83% to 53% in experimental trials when heat stress was applied during plant
46 emergence and grain fill stages (Ali et al., 2019). In Tasmania, wheat yields are projected to increase,
47 particularly at sites presently temperature-limited (Phelan et al., 2014).

48

49 New Zealand evidence on impacts across crops is very limited. Considering precipitation and temperature
50 changes alone show minor effects on crop yield, and winter yields of some crops may increase (e.g. wheat,
51 maize) (Ausseil et al., 2019b). For temperate fruit, loss of winter chill may reduce yields in some regions and
52 trigger impacts across supply chains (Cradock-Henry et al., 2019) (11.5.1). Increased pathogens could
53 damage the cut flower, guava and feijoa fruit growing, and the honey industries (Lawrence et al., 2016). The
54 combined effects of changes in seasonality, temperature, precipitation, water availability and extremes, such
55 as drought, have the potential to escalate impacts, but understanding of these effects is limited.

56

Do Not Cite, Quote or Distribute  11-47       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Other climate-change related factors complicate crop climate responses. When CO2 was elevated from
 2 present-day levels of 400 ppm to 550 ppm in trials, yields of rainfed wheat, field pea and lentil increased
 3 approximately 25% (0-70%). However, there was a 6% reduction in wheat protein that could not be offset
 4 by additional nitrogen fertilizer (O'Leary et al., 2015; Fitzgerald et al., 2016; Tausz et al., 2017). Elevated
 5 CO2 will worsen some pest and disease pressures, e.g. Barley Yellow Dwarf Virus impacts on wheat
 6 (Trbicki et al., 2015). Warmer temperatures are also expanding the potential range of the Queensland fruit
 7 fly, including into New Zealand (Aguilar et al., 2015a) threatening the horticulture industry (Sultana et al.,
 8 2017; Sultana et al., 2020). Some crop pests (e.g. the oat aphid) are projected to be negatively affected by
 9 climate change (Macfadyen et al., 2018), but so too are beneficial insects. There is large uncertainty in
10 rainfall and crop projections for northern Australia (Table 11.3). For sugarcane, impact assessment for CO2
11 at 734ppm using the A2 emission scenario at Ayr in Queensland projected modest yield increases (Singels et
12 al., 2014). Climate change are projected to adversely impact tropical fruit crops such as mangoes through
13 higher minimum and maximum temperatures reducing the number of inductive days for flowering (Clonan et
14 al., 2020).

15

16 Climate change is projected to shift agro-ecological zones (Lenoir and Svenning, 2015; Scheffers et al.,
17 2016) (high confidence). This includes the climatically determined cropping strip bounded by the inner arid
18 rangelands and the wetter coast or mountain ranges in mainland Australia (Nidumolu et al., 2012; Eagles et
19 al., 2014; Tozer et al., 2014). A narrowing of grain growing regions is projected with a shift of the inner
20 margin towards the coast under drier and warmer conditions (Nidumolu et al., 2012; Fletcher et al., 2020).
21 The economic impact of the shift depends on adaptation (Sanderson et al., 2015; Hunt et al., 2019) and how
22 resources, support industries, infrastructure and settlements adapt. Shifts in agro-ecological zones present
23 some opportunities, for example, warming is projected to be beneficial for wine production in Tasmania
24 (Harris et al., 2020).

25

26 11.3.4.1.3 Adaptation
27 Some farmers are adapting to drier and warmer conditions through more effective capture of non-growing
28 season rainfall (e.g. stubble retention to store soil water), improved water use efficiency, and matching
29 sowing times and cultivars to the environment (Kirkegaard and Hunt, 2011; Fitzer et al., 2019; Haensch et
30 al., 2021) (high confidence). Observed adaptations include new technologies that improve resource
31 efficiencies, professional knowledge and skills development, new farmer and community networks, and
32 diversification of business and household income (Ghahramani et al., 2015; De et al., 2016). For Australian
33 wheat, earlier sowing and longer season cultivars may increase yield by 2-4% by 2050, with a range of -7 to
34 +2% by 2090 (Wang et al., 2018a). In the wheat industry, breeding for improved reproductive frost tolerance
35 remains a priority (Lobell et al., 2015). Modelling suggests that, since 1990, farm management has held
36 Australian wheat yields constant, but declining rainfall and increasing temperature may have contributed to a
37 27% decline in simulated potential Australian wheat yield (Hochman et al., 2017).

38

39 Other observed incremental adaptations include later pruning in the grape industry to spread harvest period
40 and partially restore wine balance, with neutral effects on yield and cost (Moran et al., 2019; Ausseil et al.,
41 2021). The cotton sector increasingly requires shifts in sowing dates to avoid financial impacts (Luo et al.,
42 2017). During years of low water availability, rice growers have been trading water and/or shifting to dry
43 land farming (Mushtaq, 2016).

44

45 Growers in New Zealand are changing the timing of their operations, growing crops within covered
46 enclosures, and purchasing insurance (Cradock-Henry and McKusker, 2015)(Teixeira et al. 2018).
47 Investment of capital in irrigation infrastructure has increased (Cradock-Henry et al., 2018a), although its
48 effectiveness as an adaptation depends on water availability (Box 11.5). In industries based on long-lived
49 plants, such as the kiwifruit and grape industries, many of the adaptations (e.g. breeding and growing heat-
50 adapted and disease-resistant varieties) have long-lead times and require greater investment than in the
51 cropping sector (Cradock-Henry et al., 2020a). While breeding programmes for traits with enhanced
52 resilience to future climates are beginning, there is little evidence of strategic industry planning (Cradock-
53 Henry et al., 2018a).

54

55 For drought management, balancing near-term needs with long-term adaptation to increasing aridity is
56 essential (Downing et al., 2016). Insufficient and maladaptive decisions can have far-reaching effects,
57 including changes to resources, infrastructure, services and supply chains to which others have to adapt

Do Not Cite, Quote or Distribute  11-48       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 (Fleming et al., 2015; Graham et al., 2018). While there is potential for greater proportion of agriculture to
 2 be located to northern Australia, there are significant and complex agronomic, environmental, institutional,
 3 financial and social challenges for successful transformation including the risk of disruption (Jakku et al.,
 4 2016) (medium confidence).

 5

 6 11.3.4.2 Livestock

 7

 8 11.3.4.2.1 Observed impacts
 9 Both the seasonality and annual production of pasture is changing (high confidence). In many regions,
10 warming is increasing winter pasture growth (Lieffering, 2016); effects on spring growth are more mixed
11 with some regions experiencing increased growth {(Newton et al. 2014)} and others experiencing reduced
12 spring growth (Perera et al., 2020). Droughts are causing economic damage to livestock enterprises with
13 drought and market prices significantly affecting profit (Hughes et al., 2019a), in addition to the impacts on
14 animal health and the livelihoods of pastoralists, periods of drought contribute to land degradation,
15 particularly in the cattle regions of northern Australia (Marshall, 2015). Heat load in cattle leads to reduced
16 growth rates and reproduction, and extreme heat waves can lead to death (Lees et al., 2019; Harrington,
17 2020). Temperatures over 32oC reduce ewe and ram fertility along with the birth weight of lambs (van
18 Wettere et al., 2021).

19

20 11.3.4.2.2 Projected impacts
21 Some areas may experience increased pasture growth, but others may experience a decrease that cannot be
22 fully offset by adaptation (Moore and Ghahramani, 2013; Lieffering, 2016; Kalaugher et al., 2017) (high
23 confidence). Climate change may modify the seasonality of pasture growth rates more than annual yields in
24 New Zealand (Lieffering, 2016). In eastern parts of Queensland, climate change impacts on pasture growth
25 are equivocal, with simple empirical models suggesting a decrease in net primary productivity (Liu et al.,
26 2017), whilst mechanistic models that include increases in length of the growing season and the beneficial
27 effects of CO2 fertilisation indicate increases in pasture growth (Cobon et al., 2020). In Tasmania, annual
28 pasture production is projected to increase by 13­16%, even with summer growth projected to reduce with
29 increased inter-annual variability, resulting in projected increase of milk yields by 3­16% per annum (Phelan
30 et al., 2015).

31

32 Extreme climatic events (droughts, floods and heatwaves) are projected to adversely impact productivity for
33 livestock systems (medium confidence). This includes reduced pasture growth rates between 3-23% by 2070
34 from late spring to autumn, and elevated growth in winter and early spring (Cullen et al., 2009; Hennessy et
35 al., 2016; Chang-Fung-Martel et al., 2017). Heavy rainfall and storms are projected to lead to increased
36 erosion, particularly in extensively grazed systems on steeper land, reducing productivity for decades,
37 reducing soil carbon (Orwin et al., 2015), and increasing sedimentation. Increased heat stress in livestock is
38 projected to decrease milk production and livestock reproduction rates (high confidence) (Nidumolu et al.,
39 2014; Ausseil et al., 2019b; Lees et al., 2019). In Australia, the average number of moderate to severe heat
40 stress days for livestock is projected to increase 12-15 days by 2025 and 31­42 days by 2050 compared to
41 1970-2000 (Nidumolu et al., 2014). In New Zealand, an extra 5 (RCP2.6) to 7 (RCP8.5) moderate heat stress
42 days per year are projected for 2046-2060 (Ausseil et al., 2019b) (high confidence) especially affecting
43 animals transported long distances (Zhang and Phillips, 2019) and strain the cold chains needed to deliver
44 meat and dairy products safely. The distribution of existing and new pests and diseases are projected to
45 increase, for example, new tick and mosquito-borne diseases such as Bovine ephemeral fever (Kean et al.,
46 2015).

47

48 11.3.4.2.3 Adaptation
49 Adaptations in both grazing and confined beef cattle systems require enhanced decision-making skills
50 capable of integrating biophysical, social and economic considerations (high confidence). Social learning
51 networks that support integration of lessons learned from early adopters and involvement with science-based
52 organizations can help enhance decision-making and climate adaptation planning (Derner et al., 2018).
53 Pasture management adaptations for livestock production include deeper rooted pasture species in higher
54 rainfall regions (Cullen et al., 2014) and drought tolerant species (Mathew et al., 2018). Soil and land
55 management practices are important in ensuring soils maintain their supporting and regulating services
56 (Orwin et al., 2015). Adaptations in the primary sector in New Zealand are now positioned within the
57 requirements of the National Policy Statement on Freshwater (MfE, 2020b). Adaptations to manage heat

Do Not Cite, Quote or Distribute  11-49       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 stress in livestock include altering the breeding calendar, providing shade and sprinklers, altering nutrition
 2 and feeding times, and more heat-tolerant animal breeds (Chang-Fung-Martel et al., 2017; Lees et al., 2019;
 3 van Wettere et al., 2021).

 4

 5 Beef rangeland systems in Queensland are projected to have benefits in the south-east through higher CO2
 6 and temperatures extending the growing season and reducing frost, but a warmer and drier climate in the
 7 south-west may reduce pasture and livestock production (Cobon et al., 2020). Northern Queensland is most
 8 resilient to temperature and rainfall changes (production limited by soil fertility) while western/central west
 9 Queensland is most sensitive to rainfall changes. i.e. low rainfall associated with lower productivity (Cobon
10 et al., 2020). The social context of climate change impacts and the processes shaping vulnerability and
11 adaptation, especially at the scale of the individual, are critical to successful adaptation efforts.(Marshall and
12 Stokes, 2014)

13

14 11.3.4.3 Forestry

15

16 11.3.4.3.1 Observed impacts
17 Climate change may have increased tree mortality in Australia's commercial Eucalyptus globulus and Pinus
18 radiata plantation forests (Crous et al., 2013; Pinkard et al., 2014). Climate warming decreased fine root
19 biomass of E. globulus (Quentin et al., 2015) and enhanced tree water use and vulnerability to heat (Crous et
20 al., 2013). Increases in fire frequency and intensity in forests of southern Australia are leading to diminishing
21 resources available for timber production (Pinkard et al., 2014) [Box 11.1].

22

23 11.3.4.3.2 Projected impacts
24 The projected declines in rainfall in far southwest and far southeast mainland Australia are projected to
25 reduce plantation forest yields (high confidence). Warmer temperatures are projected to reduce forest growth
26 in hotter regions (between 7-25%), especially where species are grown at the upper range of their
27 temperature tolerances, and increase plantation forest growth (>15%) in cooler margins like Tasmania and
28 the Victorian highlands (2030, A2); emission scenario A2 creates a warming trajectory slightly higher than
29 the RCP6.0 warming scenario, but less than RCP8.5 (Rogelj et al., 2012; Battaglia and Bruce, 2017).
30 Elevated CO2 is projected to increase forest growth if other biophysical factors are not limiting (medium
31 confidence) (Quentin et al., 2015; Duan et al., 2018).

32

33 Forestry plantations are projected to be negatively impacted from increases in fire weather (Box 11.1),
34 particularly in southern Australia (Pinkard et al., 2014) (high confidence). Increased pest damage due to
35 temperature increases may reduce eucalypt and pine plantation growth by as much as 40% in some
36 Australian environments by 2050) (Pinkard et al., 2014). Increased heat and water stress may enhance insect
37 pest defoliation for P. radiata in Australia (e.g. Sirex noctilio, Ips grandicollis and Essigella californica)
38 (Mead, 2013; Pinkard et al., 2014).

39

40 Combined impacts from heavy rainfall, soil erosion, drought, fire and pest incursions are projected to
41 increase risks to the permanence of carbon offset and removal strategies in New Zealand for meeting its
42 climate change targets (PCE, 2019; Watt et al., 2019; Anderegg et al., 2020; Schenuit et al., 2021). Effective
43 management of the interactions between mitigation and adaptation policies can be achieved through
44 governance and institutions, including Mori tribal organisations and sectoral adaptation, to ensure effective
45 and continued carbon sequestration and storage as the climate changes (Lawrence et al., 2020b) (11.4.2)
46 (Box 11.5) (medium confidence). The productivity of radiata pine (P. radiata D. Don) in New Zealand due to
47 higher CO2 is projected to increase by 19% by 2040 and 37% by 2090, but greater wind damage to trees is
48 expected (Watt et al., 2019). Changes in the distribution of existing weeds, pests and diseases with potential
49 establishment of new subtropical pests and seasonal invasions are projected (Kean et al., 2015; Watt et al.,
50 2019; MfE, 2020a). Increased pathogens such as pitch canker, red needle cast and North American bark
51 beetles could damage plantations (Hauraki Gulf Forum, 2017; Lantschner, 2017; Watt et al., 2019).

52

53 11.3.4.3.3 Adaptation
54 Adaptation options include: increased investment in monitoring forest condition and functioning; early
55 detection and management of insect pests, diseases and invasive species; improved selection of land with
56 appropriate growing conditions for plantation timber production under current and future conditions; trialling
57 new species and genetic varieties; changing timing and frequency of planned fuel reduction fires, introducing

Do Not Cite, Quote or Distribute  11-50       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 more fire-tolerant tree species where appropriate, reducing ignition sources and maintaining access and
 2 emergency response capacity (Boulter, 2012; Pinkard et al., 2014; Keenan, 2017).

 3

 4 11.3.4.4 Marine Food

 5

 6 11.3.4.4.1 Observed impacts
 7 Ecological impacts of climate change on fisheries species have already emerged (Morrongiello and Thresher,
 8 2015; Gervais et al., 2021)(high confidence). This includes loss of habitats for fisheries species (Vergés et
 9 al., 2016; Babcock et al., 2019), and poleward shifts in distribution of barrens-forming urchins (Ling and
10 Keane, 2018) impacting abalone and rock lobster fisheries. The percentage of reef as barrens across eastern
11 Tasmania grew from 3.4% to 15.2% from 2001/02 to 2016/17, a ~10.5% increase per annum over the 15-
12 year period (Ling and Keane, 2018). Oysters farmed from wild spat (Sydney rock oysters Saccostrea
13 glomerata) are most at risk from climate change, primarily due to observed increases in summer
14 temperatures and heat wave-related mortalities (Doubleday et al., 2013). The exceptional 2017/18 summer
15 heatwave caused significant losses of farmed salmon in New Zealand, with farm owners seeking consent to
16 move operations to cooler water (Salinger et al., 2019b).

17

18 11.3.4.4.2 Projected impacts
19 Aquaculture is projected to be more easily adapted than wild fisheries to avoid excessive exposure to the
20 physio-chemical stresses from acidification, warming and extreme events (Richards et al., 2015). In New
21 Zealand, wild and cultured shellfish are identified as most at risk from climate change (Capson and Guinotte,
22 2014). Changes in ocean temperature and acidification, and the downstream impacts on species distribution,
23 productivity and catch are projected concerns (Law et al., 2016) (medium confidence) that impact Mori
24 harvesting of traditional seafood, and the social, cultural and educational elements of food gathering
25 (mahinga kai) (MfE, 2016). Warm temperate hatchery-based finfish species (yellowtail kingfish Seriola
26 lalandi) are projected to be the least at risk, because of well controlled environmental conditions in
27 hatcheries, and temperature increases which are expected to increase growth rates and productivity during
28 the grow-out stage (Doubleday et al., 2013). For wild fisheries, multi-model projections suggest temperate
29 and demersal systems, especially invertebrate shallow water species, would be more strongly affected by
30 climate change than tropical and pelagic systems (Pecl et al., 2014; Fulton et al., 2018; Pethybridge et al.,
31 2020) (medium confidence). In New Zealand waters, available habitat for both albacore tuna and oceanic
32 tuna (Cummings et al., 2021) is expected to widen and shift.

33

34 11.3.4.4.3 Adaptation
35 Selective breeding in oysters is projected to be an important global adaptation strategy for sustainable
36 shellfish aquaculture which can withstand future climate-driven change to habitat acidification (Fitzer et al.,
37 2019). Less than a quarter of fisheries management plans for 99 of Australia's most important fisheries
38 considered climate change, and only to a limited degree (Fogarty et al., 2019; Fogarty et al., 2021).
39 Implementation of management and policy responses to climate change have lagged in part because climate
40 change has not been considered as the most pressing issue (Hobday and Cvitanovic, 2017; Fogarty et al.,
41 2019; Fogarty et al., 2021) (Cross-Chapter Box MOVING SPECIES in Chapter 5).

42

43

44 [START BOX 11.5 HERE]

45

46 Box 11.5: New Zealand's Land, Water and People Nexus under a changing climate

47

48 New Zealand's economy, dominated by the primary sector and the tourist industry (pre-COVID), relies upon
49 a "clean green" image of water, natural ecosystems and pristine landscapes (Foote et al., 2015; Roche and
50 Argent, 2015; Hayes and Lovelock, 2017). Water is highly valued by Mori for its mauri or life force and for
51 its intrinsic values and multiple uses (Harmsworth et al., 2016). Increasingly, these diverse values are in
52 conflict (Hopkins et al., 2015) due to increasing pressures from how land is used and managed and the
53 effects on water availability and quality. Such tensions will be further challenged as temperatures rise and
54 extreme events intensify beyond what has been experienced, thus stressing current adaptive capacities
55 (Hughey and Becken, 2014; Cradock-Henry and McKusker, 2015; Hopkins et al., 2015; MfE and StatsNz,
56 2021) (11.2.2; 11.3.4) (high confidence).

57

Do Not Cite, Quote or Distribute  11-51       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Irrigation has increasingly been used to enhance primary sector productivity and regional economic
 2 development (Srinivasan et al., 2017; Fielke and Srinivasan, 2018; MfE and StatsNz, 2021). Pressure for
 3 long-term access to groundwater or large-scale water storage is increasing to ensure the ongoing viability of
 4 the primary sector as the climate changes. While investment in irrigation infrastructure may reduce climate
 5 change impacts in the short-term, maladaptive outcomes cannot be ruled out longer-term which means that
 6 focusing attention now on adaptive and transformational measures can help increase climate resilience in
 7 areas exposed to increasing drought and climate extremes that disrupt production (Abel et al., 2016;
 8 Cradock-Henry et al., 2019) (Yletyinen et al., 2019) (medium confidence).

 9

10 Furthermore, over-allocation raises further tensions from competing uses of water such as for horticulture
11 and urban water supplies, as well as for ecological requirements. The deterioration of water quality and loss
12 of places of social, economic, cultural, and spiritual significance creates increasing tension for Mori
13 especially (Harmsworth et al., 2016; Salmon, 2019; MfE and StatsNz, 2021). Public concern has increased
14 about the deterioration of New Zealand's waterways and the profiting of some land uses at the expense of
15 environmental quality and human health - tensions that make adaptation to climate change more challenging
16 (Duncan, 2014; Foote et al., 2015; Scarsbrook and Melland, 2015; McDowell et al., 2016; McKergow et al.,
17 2016; Greenhalgh and Samarasinghe, 2018). A lack of precautionary governance of water resources linked to
18 unsustainable land use practices degrading water quality (Scarsbrook and Melland, 2015; Salmon, 2019)
19 highlights the role that foresight could play in managing the nexus between land, water and people in a
20 changing climate (11.3.3). Adaptive planning has potential for navigating these multi-dimensional challenges
21 (Sharma-Wallace et al., 2018; Cradock-Henry and Fountain, 2019; Hurlbert et al., 2019) (11.7).

22

23 Furthermore, land and particularly plantation and native forests play a critical role in meeting New Zealand's
24 emissions reduction goals. However, the persistence of land and forests as a carbon sink is uncertain and the
25 sequestered carbon is at risk from future loss resulting from climate change impacts, including from
26 increased fire, drought and pest incursions, storms and wind (IPCC, 2019a; PCE, 2019; Watt et al., 2019;
27 Anderegg et al., 2020) (11.3.4.3), emphasising the importance of interactions between mitigation and
28 adaptation policy and implementation. Integrated climate change policies across biodiversity, water quality,
29 water availability, land use and forestry for mitigation can support the management of land use, water and
30 people conflicts, but there is little evidence of such coordinated policies (Cradock-Henry et al., 2018b;
31 Wreford et al., 2019). Implementation of the National Policy Statement for Freshwater Management 2020
32 (MfE, 2020b) and the National Adaptation Plan (due August 2022) present opportunities for such
33 interconnections and diverse values to be addressed, as well as enabling sector and community benefits to be
34 realised across New Zealand (Awatere et al., 2018; Lawrence et al., 2020b).

35

36 [END BOX 11.5 HERE]

37

38

39 11.3.5 Cities, Settlements and Infrastructure

40

41 Almost 90% of the population of Australia and New Zealand is urban (World Bank, 2019). Each country has
42 vibrant and diverse urban, rural and remote settlements, with some highly disadvantaged areas isolated by
43 distance and limited infrastructure and services (Argent et al., 2014; Charles-Edwards et al., 2018; Spector et
44 al., 2019). Some areas in northern Australia and New Zealand, especially those with higher proportions of
45 Indigenous inhabitants, face severe housing, health, education, employment and services issues (Kotey,
46 2015) which increases their vulnerability to climate change.

47

48 Infrastructure within and between cities and settlements is critical for activity across all sectors, with
49 interdependencies increasing exposure to climate hazards (11.5.1). Previous planning horizons for existing
50 infrastructure are compromised by now having to accommodate ongoing sea-level rise, warming, and
51 increasing frequency of extreme rainfall and storm events (Climate Institute, 2012; MfE, 2017a). There is
52 almost no information on the costs and benefits of adapting vulnerable and exposed infrastructure in
53 Australia or New Zealand. Given the value of the infrastructure and the rising damage costs, this represents a
54 large knowledge gap leading to an adaptation investment deficit.

55

56 11.3.5.1 Observed Impacts

57

Do Not Cite, Quote or Distribute  11-52       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Critical infrastructure, cities and settlements are being increasingly affected by chronic and acute climate
 2 hazards including heat, drought, fire, pluvial and fluvial flooding and sea-level rise, with consequent effects
 3 for many sectors (Instone et al., 2014; Loughnan et al., 2015; Zografos et al., 2016; Hughes et al., 2021)
 4 (high confidence). Risks and impacts vary with physical characteristics, location, connectivity and socio-
 5 economic status of settlements because of the ways these influence exposure and vulnerability (Loughnan et
 6 al., 2013; MfE, 2020a) (high confidence).

 7

 8 Weather-related disasters are causing significant disruption and damage (Paulik et al., 2019a; CSIRO, 2020;
 9 Paulik et al., 2020). In Australia, during 1987-2016, natural disasters caused an estimated 971 deaths and
10 4,370 injuries, 24,120 people were made homeless and about 9 million people were affected (Deloitte,
11 2017a). More than 50% of these deaths and injuries came from heatwaves in cities and 22% from fires.
12 During 2007-2016, Australia natural disaster costs averaged A$18.2 billion per year with largest
13 contributions from floods (A$8.8 billion), followed by cyclones (A$3.1 billion), hail (A$2.9 billion), storms
14 (A$2.3 billion) and fires (A$1.1 billion) (Deloitte, 2017a). The Australian fires in 2019-2020 cost over A$8
15 billion, with devastating impacts on settlements and infrastructure (Box 11.1)

16

17 Sea-level rise affects many interdependent systems in cities and settlements which increase the potential for
18 compounding and cascading impacts (11.5.1). Seaports, airports, water treatment plants, desalination plants,
19 roads and railways are increasingly exposed to sea-level rise (very high confidence), impacting their
20 longevity, levels of service and maintenance (high confidence) (McEvoy and Mullett, 2014; Woodroffe et
21 al., 2014; PCE, 2015; Ranasinghe, 2016; Newton et al., 2018; Paulik et al., 2020) (Box 11.6). Compounding
22 coastal hazards in New Zealand, such as elevated water tables associated with rising sea-level and intense
23 rainfall (Morgan and Werner, 2015; McBride et al., 2016; White et al., 2017; Hughes et al., 2021) are
24 exerting pressure on stormwater and wastewater infrastructure and drinking water supply and quality (MfE,
25 2020a).

26

27 Extreme heat events exacerbate problems for vulnerable people and infrastructure in urban Australia where
28 urban heat is superimposed upon regional warming, and there are adverse impacts for population and
29 vegetation health, particularly for socio-economically disadvantaged groups (Tapper et al., 2014; Heaviside
30 et al., 2017; Filho et al., 2018; Gebert et al., 2018; Rogers et al., 2018; Longden, 2019; Marchionni et al.,
31 2019; Tapper, In Press) (11.3.6), energy demand, energy supply and infrastructure (Newton et al., 2018)
32 (11.3.10) (very high confidence). Extreme heat is increasingly threatening liveability in some rural areas in
33 Australia (Turton, 2017), particularly given their reliance on outside physical work and older populations.
34 Settlement design and the level of greening interact with climate change to influence local heating levels
35 (Tapper et al., 2014; Wong et al., 2020; Tapper, In Press).

36

37 Floods cause major damage. The floods of early 2019 in north Queensland cost A$5.68 billion (Deloitte,
38 2019), while Cyclone Yasi and the Queensland floods of 2011 cost A$6.9 billion (Deloitte, 2016).
39 Floodplains in New Zealand have considerably higher overall national exposure of buildings and population
40 than coasts (Paulik et al., 2019a) (Box 11.4). The insured loss from the 12 costliest floods in New Zealand
41 from 2007-2017 totalled NZ$471.56 million, of which NZ$140.48 million could be attributed to climate
42 change (Frame et al., 2020).

43

44 Climatic extremes are exacerbating existing vulnerabilities (high confidence). Long supply chains, poorly
45 maintained infrastructure, social disadvantage and poor health, and lack of skilled workers (Eldridge and
46 Beecham, 2018; Mathew et al., 2018; Rolfe et al., 2020) are contributing to serious stress and disruption
47 (Smith and Lawrence, 2014; Kiem et al., 2016). In many rural settlements, population ageing and reliance on
48 an over-stretched volunteer base for recovery from extreme events are increasing vulnerability to climate
49 change (Astill and Miller, 2018; Davies et al., 2018). Recovery from long, intense, more frequent and
50 compounding climatic events in rural areas has been disrupted by the erosion of natural, financial, built,
51 human and social capital (De et al., 2016; Sheng and Xu, 2019). Delayed recovery from extreme climatic
52 events has been compounded by long-term displacement which in turn prolongs the impacts (Matthews et
53 al., 2019). Severe droughts have contributed to poor health outcomes for rural communities, including
54 extreme stress and suicide (Beautrais, 2018; Perceval et al., 2019). In Australia, competition between water
55 users has left some rural communities experiencing extreme water shortage and insecurity with associated
56 health impacts (Wheeler et al., 2018; Judd, 2019) (Box 11.3).

57

Do Not Cite, Quote or Distribute  11-53       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 11.3.5.2 Projected Impacts

 2

 3 Changes in heat waves, droughts, fire weather, heavy rainfall, storms and sea-level rise are projected to
 4 increase negative impacts for cities, settlements and infrastructure (Tables 11.3a and 11.3b; Boxes 11.1,
 5 11.3, 11.4) (high confidence).

 6

 7 Increased floods, coastal inundation (assuming a sea-level rise of 1.6 m by 2100), wildfires, windstorms and
 8 heatwaves may cause property damage in Australia estimated at A$91 billion per year by 2050 and A$117
 9 billion per year by 2100 for RCP8.5, while damage-related loss of property value is estimated at A$611
10 billion by 2050 and A$770 billion by 2100 for RCP8.5 (Steffen et al., 2019). For 1.0 m sea-level rise, the
11 value of exposed assets in New Zealand would be NZ$25.5 billion (Box 11.6). For 1.1 m sea-level rise, the
12 value of exposed assets in Australia would be A$164-226 billion (Box 11.6). These cost estimates exclude
13 impacts on personal livelihood, well-being or lifestyle.

14

15 Extreme heat risks are projected to exacerbate existing heat-related impacts on human health, vegetation and
16 infrastructure (Tapper et al., 2014; Tapper, In Press) (11.3.6). In Australia, the annual frequency of days over
17 35oC is projected to increase 20-70% by 2030 (RCP4.5), and 25­85% (RCP2.6) to 80­350% (RCP8.5) by
18 2090 (Table 11.3a). For example, Perth may average 36 days over 35oC by 2030 (RCP4.5). In New Zealand,
19 the annual frequency of days over 25oC may increase 20-60% (RCP2.6) to 50-100% (RCP8.5) by 2040, and
20 20­60% (RCP2.6) to 130-350% (RCP8.5) by 2090 (Table 11.3b). For example, Auckland may average 39
21 days over 25oC by 2040 (RCP8.5). Unprecedented extreme temperatures, as high as 50oC in Sydney or
22 Melbourne, could occur with global warming of 2.0oC (Lewis et al., 2017). Heat-related costs for Melbourne
23 during 2012-2051 are estimated at A$1.9 billion, of which A$1.6 billion is human health/mortality costs
24 (AECOM, 2012). Extreme heat is threatening liveability in some rural areas in Australia (Turton, 2017),
25 particularly given their reliance on outside physical work and older populations.

26

27 Key infrastructure and services face major challenges. Structural metal corrosion rates are projected to
28 increase significantly at coastal locations but decrease inland (Trivedi et al., 2014). A drier climate may
29 decrease the rate of deterioration of road pavements but extreme rainfall events and heat pose a significant
30 risk (Taylor and Philp, 2015), especially to unsealed roads in northern Australia (CoA, 2015). Critical
31 infrastructure on coasts is at risk from sea-level rise and storm surges (Box 11.6). Facilities such as hospitals
32 face weather-related hazards exacerbated by climate change and not originally anticipated in building and
33 infrastructure design (Loosemore et al., 2011; Loosemore et al., 2014). By 2050, increased risks are
34 projected for the availability and quality of potable water supplies, delivery of wastewater and stormwater
35 services to communities, transport systems, electricity infrastructure, operating municipal landfills, and
36 contaminated sites located near rivers and the coast (Gilpin et al., 2020; MfE, 2020a; Hughes et al., 2021).
37 These then create risks to social cohesion and community wellbeing from displacement of individuals,
38 families and communities, with inequitable outcomes for vulnerable groups (Boston and Lawrence, 2018).

39

40 11.3.5.3 Adaptation

41

42 In cities and settlements, climate adaptation is underway and is being led and facilitated by state and local
43 government leadership and facilitation, particularly in Australia (Hintz et al., 2018; Newton et al., 2018)
44 (Table 11.7, Supplementary Material Table SM11.1a) (high confidence).

45

46 Effective adaptations to urban heat include spatial planning, expanding tree canopy and greenery, shading,
47 sprays and heat-resistant and energy-efficient building design, including cool materials and reflective or
48 green roofs (very high confidence) (Broadbent et al., 2018; Jacobs et al., 2018b; Haddad et al., 2019; Haddad
49 et al., 2020a; Yenneti et al., 2020; Bartesaghi-Koc et al., 2021; Tapper, In Press). Reducing urban heat not
50 only benefits human health but reduces demand for, and cost of, air conditioning (Haddad et al., 2020b) and
51 the risk of electricity blackouts (11.3.10).

52

53 Adaptation progress is being hampered by current urban redevelopment practice and statutory planning
54 guidelines that are leading to removal of critical urban green space (Newton and Rogers, 2020). Reform of
55 approaches to urban redevelopment would facilitate adaptation (Newton and Rogers, 2020). Several cities in
56 Australia and New Zealand are part of the 100 Resilient Cities global network which helped facilitate the
57 metropolitan Melbourne Urban Forest Strategy across councils (Fastenrath et al., 2019; Coenen et al., 2020)

Do Not Cite, Quote or Distribute  11-54       Total pages: 151
    FINAL DRAFT                                Chapter 11                          IPCC WGII Sixth Assessment Report

1 and in New Zealand, restoration of the urban forest in Hamilton is reducing heat stressors (Wallace and
2 Clarkson, 2019). In peri-urban zones, adapting to fire risk is a contested issue, raising difficult trade-offs
3 between heat management, ecological values and fuel reduction in treed landscapes (Robinson et al., 2018).

 4

 5 The resilience of Australia's major cities to flooding and drought has been advanced through a range of
 6 economic and physical interventions. Water sensitive urban design irrigates vegetation with harvested storm
 7 water that improves water security, flood risk, carbon sequestration, biodiversity, air and water quality, and
 8 delivers cooling that can save human lives in heatwaves (Wong et al., 2020). Storm water harvesting is
 9 supported by some councils in New Zealand and can deliver recycled water for households (Attwater and
10 Derry, 2017), improving climate resilience and reducing water demand (White et al., 2017). Addressing
11 infrastructure vulnerability is essential given the long lifetime of the assets, criticality of services and high
12 costs of maintenance (Chester et al., 2020; Hughes et al., 2021).

13

14 Climate risk management is developing, but adaptive capacity, implementation, monitoring and evaluation
15 are uneven across all scales of cities, settlements and infrastructure (very high confidence) (Tables 11.15a
16 and 11.15b; Supplementary Material Tables SM11.1a, and SM11.1b). There is increasing awareness of the
17 need to move from incremental coping and defensive coastal strategies (Jongejan et al., 2016) to
18 transformational adaptation, for example, managed retreat (Torabi et al., 2018; Hanna, 2019), and to consider
19 the flow-on effects (e.g. for housing and employment) (Fatori et al., 2017; Torabi et al., 2018). Strategies
20 limited to building household and community self-reliance (Astill and Miller, 2018) are increasingly
21 inadequate given systemic and interconnected stressors and cascading impacts across interdependent systems
22 (Lawrence et al., 2020b). Integrated approaches to climate change adaptation and emissions reduction have
23 potential for addressing interdependent systems (e.g. nature-based approaches, climate-sensitive urban
24 design, energy and transport systems) (Norman et al., 2021). Climate risk assessment and adaptation
25 guidelines have been prepared for transport infrastructure authorities and organisations (Finlayson et al.,
26 2017; Byett et al., 2019; Yenneti et al., 2020).

27

28

29 Table 11.7: Cities, settlements and infrastructure: key risks and adaptation options.

    Sector           Key Risks                 Adaptation Options     Inter-Sector              Sources
                                                                      Dependencies

    Road             Heat; sea-level rise;     Re-routing; coastal    Ports (fuel supply); rail (NCCARF, 2013;
                     coastal surges; floods    protection; improved
                     and high intensity        drainage               (fuel supply);            CoA, 2018a; MfE,
                     rainfall impacts on
                     road foundations                                 electricity               2020a)

    Rail             Extreme temperatures; Drainage and               Electricity;              (CoA, 2018a; MfE,

                     flooding; sea-level ventilation                  telecommunications; 2020a)

                     rise; high intensity      improvements;          fuel supply (transport,

                     rainfall impacts on systematic risk              ports)

                     track foundations         assessments; overhead

                                               wire and rail/sleeper

                                               upgrades; rerouting

    Urban and Rural  Extreme temperatures;     Multiple options from  Road; rail; electricity;  (CoA, 2018a; Newton
                     floods; extreme           the building-to-city   air and seaports;         et al., 2018; Haddad et
    Built            weather events;           scale to reduce heat   telecommunications;       al., 2019; MfE, 2020a;
    Environment1     wildfire (at urban-rural  impacts and improve    water and wastewater      Paulik et al., 2020;
                     interface); sea-level     climate resilience;                              Tapper, In Press)
                     rise                      behavioural change;                              (Box 11.4)
                                               coastal defences and                             (Box 11.4)
                                               managed retreat

    Do Not Cite, Quote or Distribute           11-55                                            Total pages: 151
FINAL DRAFT                               Chapter 11                      IPCC WGII Sixth Assessment Report

Electricity     High wind/                Demand management;       Road; rail; water          (CoA, 2017; MfE,
                temperature events;       re-engineering and                                  2020a)
                wildfire; lightning;      new technology;                                     (11.3.10.)
                dust storms; drought      network intelligence;
                (hydro)                   smart metering;
                                          improved planning for
                                          outages

Ports: Air and  Sea-level rise; coastal   Air; improved coastal, Electricity; road; rail,     (McEvoy and Mullett,
Sea             surges; wind; heat;                                                           2014; MfE, 2020a)
                extreme weather           pluvial and fluvial      water
                events
                                          flood protection, on-

                                          site services. Sea;

                                          widening operational

                                          limits, raising wharfs,

                                          roads and breakwaters.

Telecommunicat Floods; wildfires;         Protect; place           Electricity; digital       (NCCARF, 2013)
                                          underground; wireless
ions            extreme wind              systems                  connectivity; all sectors

                                                                   serviced; rural

                                                                   communities

Stormwater      High intensity rainfall;  Large investments in     Electricity;               (White et al., 2017;
                increased and extreme     upgrading centralized    telecommunications;        CoA, 2018a; Gilpin et
Wastewater and  temperatures;             infrastructure and       urban and rural built      al., 2020; MfE, 2020a;
Water supply1.  flooding; drought; sea-   capacity; increasing     environment                Wong et al., 2020;
                level rise                investment in                                       Hughes et al., 2021)
                                          decentralized                                       (Box 11.4)
                                          infrastructure and
                                          capacity (e.g. Water
                                          Sensitive Urban
                                          Design); demand
                                          management; fewer
                                          options in smaller
                                          communities;
                                          governance at scale

          Table Notes:
          1.Water supply safety and security and exposure of buildings have been identified as the most significant risks for
          New Zealand in terms of urgency and consequence (MfE, 2020a). No such ranking of risk has been done for
          Australia.

 1

 2

 3 Infrastructure service vulnerability in New Zealand is supported by new institutional adaptations including
 4 the Infrastructure Commission to develop a 30-year national infrastructure strategy. The Climate Change
 5 Commission (Climate Change Commission, 2020) has issued six principles for climate-relevant
 6 infrastructure investments and is mandated to monitor the National Climate Change Adaptation Plan based
 7 on the first National Climate Change Risk Assessment (MfE, 2020a). A National Disaster Resilience
 8 Strategy addresses integrated planning for risk reduction and awareness-raising in New Zealand (Department
 9 of the Prime Minister and Cabinet, 2019).

10

11 Successive inquiries and reviews highlight potential synergies between disaster risk management and climate
12 resilience (11.5.1) (Smith and Lawrence, 2018; Ruane, 2020). In Australia, there is a National Disaster Risk
13 Reduction Framework (CoA, 2018b) and a National Recovery and Resilience Agency (CoA, 2021) that help
14 underpin the development of national support systems for rural and regional emergency management and
15 associated volunteer sectors (McLennan et al., 2016) and wildfire smoke impacts (CoA, 2020e). The
16 National Heatwave Framework Working Group uses a Heatwave Forecast Service, and heatwave early
17 warning and adaptation systems that operate in Adelaide, Melbourne, Sydney and Brisbane have reduced
18 potential death rates (Nitschke et al., 2016).

19

20 Infrastructure planning is lagging behind international standards for climate resilience evaluation and
21 guidance for adaptation to climate risk (CSIRO, 2020; Kool et al., 2020; Hughes et al., 2021) (high

Do Not Cite, Quote or Distribute          11-56                                               Total pages: 151
    FINAL DRAFT                                       Chapter 11                IPCC WGII Sixth Assessment Report

1 confidence). Some companies have examined their exposure to climate risk and developed strategies to
2 minimise their vulnerability (Climate Institute, 2012) (11.3.8). Climate risk assessments have been
3 conducted for the electricity sector in both Australia and New Zealand (11.3.10). Climate change is
4 considered in Australian infrastructure plans for national and regional water supply security, water for
5 irrigated agriculture, a coastal hazards adaptation strategy, and the Tanami Road upgrade (Infrastructure
6 Australia, 2016; Infrastructure Australia, 2019; Infrastructure Australia, 2021).

 7

 8 Industry associations are beginning to facilitate climate adaptation for infrastructure, including the Australian
 9 Green Infrastructure Council (CoA, 2015), the Green Building Council of Australia, Green Star Programme
10 (GBCA, 2020), the Water Services Association of Australia, Climate Change Adaptation Guidelines
11 (WSAA, 2016) and the Australian Sustainable Built Environment Council, Built Environment Adaptation
12 Framework (ASBEC, 2012). The Infrastructure Sustainability Rating Scheme measures the social,
13 environmental, governance and cultural outcomes delivered by more than $160 billion worth of
14 infrastructure, and it is projected to deliver a cost-benefit ratio of 1:1.6 to 1:2.4 during 2020-2040 (RPS,
15 2020). There is scope for engagement of industry in transitioning to a low carbon green economy that is
16 adapted to climate change, but less certainty on how to develop appropriate business cases (Newton and
17 Newman, 2015).

18

19 There are tensions between settlement-scale adaptation options such as managed retreat that focus on the
20 long term, and people's values, place attachments, needs and capacities (Gorddard et al., 2016; Fatori et al.,
21 2017; Graham et al., 2018; O'Donnell, 2019; Norman et al., 2021). Tensions also exist between climate
22 change adaptation and mitigation goals (e.g. current energy efficiency standards in Australian buildings can
23 worsen their heat resistance and increase dependence on air-conditioning) (Hatvani-Kovacs et al., 2018).
24 Where there is a lack of coordination between jurisdictions, there can be flow-on effects from failure to
25 adapt, for example in coastal local government areas (Dedekorkut-Howes et al., 2020) (Box 11.6). There is
26 limited information across the region on climate change impacts and adaptation options for
27 telecommunications (NCCARF, 2013) (Table 11.7). There is an emerging recognition that implementing and
28 evaluating the adaptation process (vulnerability and risk assessments, identification of options, planning,
29 implementation, monitoring, evaluation and review) in local contexts can advance more effective adaptation
30 (Moloney and McClaren, 2018). For example, the Victorian State Government has built monitoring,
31 evaluation and adaptation components into its adaptation plan (Table 11.15a).

32

33

34 [START BOX 11.6 HERE]

35

36 Box 11.6: Rising to the Sea-Level Challenge

37

38 Many of the region's cities and settlements, cultural sites and place attachments are situated around harbours,
39 estuaries and lowland rivers (Black, 2010; PCE, 2015; Australia SoE, 2016; Rouse et al., 2017; Hanslow et
40 al., 2018; Birkett-Rees et al., 2020) exposed to ongoing relative sea-level rise (RSLR). RSLR includes
41 regional variability in oceanic conditions (Zhang et al., 2017) and vertical land movement along New
42 Zealand's tectonically dynamic coasts (Levy et al., 2020) and some Australian hotspots for subsidence
43 (Denys et al., 2020; King et al., 2020; Watson, 2020).

44

45

46 Table Box 11.6.1: Observed and projected impacts from higher mean sea level

    Impacts from increase in mean sea level                         References

    Nuisance and extreme coastal flooding have increased from       (Hunter, 2012; McInnes et al., 2016; Stephens et al.,
    higher mean sea level in New Zealand. Projected sea level rise  2017; Stephens et al., 2020) (Steffen et al., 2014;
    will cause more frequent flooding in Australia and New          PCE, 2015; MfE, 2017a; Hague et al., 2019; Paulik
    Zealand before mid-century (very high confidence)               et al., 2020)

    Squeeze in intertidal habitats (high confidence)                (Steffen et al., 2014; Peirson et al., 2015; Mills et al.,
                                                                    2016a; Mills et al., 2016b; Pettit et al., 2016; Rouse
                                                                    et al., 2017; Rayner et al., 2021)

    Do Not Cite, Quote or Distribute                  11-57                     Total pages: 151
   FINAL DRAFT                                       Chapter 11       IPCC WGII Sixth Assessment Report

   Significant property and infrastructure damage (high          (Steffen et al., 2014; PCE, 2015; Harvey, 2019;
   confidence)                                                   LGNZ, 2019; Paulik et al., 2020) (Table Box 11.5.2)
                                                                 (Table Box 11.6.2)

   Loss of significant cultural and archaeological sites and     (Bickler et al., 2013; Birkett-Rees et al., 2020; NZ
   projected to compound with several hazards over this century  Archaeological Association, 2020)
   (medium confidence)

   Increasing flood risk and water insecurity with health and well- (Steffen et al., 2014; McInnes et al., 2016;

   being impacts on Torres Strait Islanders (high confidence)    McNamara et al., 2017)

   Degradation and loss of freshwater wetlands (high confidence) (Pettit et al., 2016; Bayliss and Ligtermoet, 2018;
                                                                                Tait and Pearce, 2019; Grieger et al., 2020; Swales
                                                                                et al., 2020)

1

 2

 3 Coastal shoreline position is driven by a complex combination of natural drivers, past and present human
 4 interventions, climate variability (Bryan et al., 2008; Helman and Tomlinson, 2018; Allis and Murray Hicks,
 5 2019) and variation in sediment flux (Blue and Kench, 2017; Ford and Dickson, 2018). RSLR, to date, is a
 6 secondary factor influencing shoreline stability (medium confidence), and in Australia no definitive sea-level
 7 rise signature is yet observed in shoreline recession, nor documented in New Zealand, due to variability in
 8 shoreline position responding to storms and seasonal, annual and decadal climate drivers (Australian
 9 Government, 2009; McInnes et al., 2016; Sharples et al., 2020).

10

11 The primary impacts of rising mean sea level (Table Box 11.6.1) are being compounded by climate-related
12 changes in waves, storm surge, rising water tables, river flows and alterations in sediment delivery to the
13 coast (medium confidence). The net effect is projected to increase erosion on sedimentary coastlines and
14 flooding in low-lying coastal areas(McInnes et al., 2016; MfE, 2017a; Hanslow et al., 2018; Wu et al., 2018).
15 Waves are projected to be higher in southern Australasia and lower elsewhere (Morim et al., 2019) and storm
16 surge slightly higher in the south, slightly lower further north in New Zealand (Cagigal et al., 2019) and
17 small robust declines along southern Australia, with potentially larger changes in the Gulf of Carpentaria
18 (Colberg et al., 2019).

19

20 The cumulative direct and residual risk from RSLR and associated impacts are projected to continue for
21 centuries, necessitating on-going adaptive decisions for exposed coastal communities and assets (MfE,
22 2017c; Oppenheimer et al., 2019; Tonmoy et al., 2019) (high confidence).

23

24

25 Table Box 11.6.2: Observed relative sea-level rise (variance-weighted average) with uncertainty range (standard
26 deviation) and projected impacts on infrastructure and population of 1.1 m in Australia and 1 m in New Zealand. Sea-
27 level rise projections for 2050 and 2090 are given in Table 11.3a and Table 11.3b.

   Country Observed relative sea- Projected impacts of sea-level rise (1.1m Australia; 1.0m New

   level rise                        Zealand)

                                     Value of        Number of        Number of          Public council
                                                                      residents          assets exposed
                                     coastal urban buildings exposed  exposed

                                     infrastructure

   Do Not Cite, Quote or Distribute                      11-58                                   Total pages: 151
FINAL DRAFT                                        Chapter 11      IPCC WGII Sixth Assessment Report

Australia 2.2±1.8 mm/year to A$164 to 187,000 to 274,000 N/A                        27,000 to
                                                                                    35,000 km of
         2018 for four >75-year >226 billion residential                            roads, and 1,200
                                                                                    to 1,500 km of
         records (or an average (DCCEE, buildings, 5,800 to                         rail lines and
                                                                                    tramways
         of 0.17 m over 75        2011; Steffen 8,600 commercial                    (DCCEE, 2011)

         years). 3.4 mm/year et al., 2019) buildings, 3,700 to

         from 1993-2019           111% rise in 6,200 light

         (Watson, 2020)           inundation industrial buildings

                                  cost from (DCCEE, 2011)

                                  2020-2100

                                  (Mallon et

                                  al., 2019)

New      1.8 mm/year from         NZ$25.5 75,890 (Paulik et        105,580          4000 km
Zealand  1900-2018, 1.2                                            (Paulik et al.,  pipelines, 1440
         mm/year from 1900-       billion          al., 2020)      2020)            km roads, 101
         1960 and 2.4 mm/year                                                       km rail, 72 km
         from 1961-2018 (Bell     (Paulik et al.,                                   electricity
         and Hannah, 2019)                                                          transmission
                                  2020)                                             lines (Paulik et
                                                                                    al., 2020)
                                                                                    NZ$5 billion
                                                                                    (2018)
                                                                                    (reserves,
                                                                                    buildings, utility
                                                                                    networks,
                                                                                    roads) (LGNZ,
                                                                                    2019)

 1

 2

 3 Prevailing decision-making assumes shorelines can continue to be maintained and protected from extreme
 4 storms, flooding and erosion, even with RSLR (Lawrence et al., 2019a). Rapid coastal development has
 5 increased exposure of coastal communities and infrastructure (high confidence) (Helman and Tomlinson,
 6 2018; Paulik et al., 2020) reinforcing perceptions of safety (Gibbs, 2015; Lawrence et al., 2015) and creating
 7 barriers to retreat and nature-based adaptations (Schumacher, 2020) (very high confidence). The efficacy and
 8 increasing costs of protection and accommodation risk reduction approaches, and rebuilding after extreme
 9 events have been questioned and have limits (PCE, 2015; MfE, 2017a; Harvey, 2019; LGNZ, 2019; Paulik et
10 al., 2020; Haasnoot et al., 2021). Future shoreline erosion is often signalled by using defined coastal setback
11 lines(s) and using probabilistic approaches to signal uncertainty (Ramsay et al., 2012; Ranasinghe, 2016).

12

13 Flooding from high spring ("king") tides or storm tides during extreme weather events are raising public
14 awareness of sea-level rise (Green Cross Australia, 2012) including through media coverage (Priestley et al.,
15 2021). The use of adaptive decision tools (11.7.3.1; Table 11.17) is increasing the understanding of changing
16 coastal risk (Bendall, 2018; Lawrence et al., 2019b; Palutikof et al., 2019b) and how dynamic adaptive
17 pathways and monitoring of them can aid implementation (Stephens et al., 2018; Lawrence et al., 2020b).
18 Collaborative governance between local governments and their communities, including with Mori tribal
19 organisations, is emerging in New Zealand (OECD, 2019b) assisted by national direction (DoC NZ, 2010)
20 and guidance on adaptive planning (Table 11.15b). This shift from reactive to pre-emptive planning is better
21 suited to ongoing RSLR (Lawrence et al., 2020b).

22

23 In Australia, adaptation to sea-level rise remains uneven across jurisdictions in the absence of clear Federal
24 or State guidance, rendering Australia unprepared for flooding from sea-level rise (Dedekorkut-Howes et al.,
25 2020). Risk-averse coastal governance at the local level has led to shifts in liabilities to other actors and to
26 future generations (Jozaei et al., 2020). Managed retreat has emerged as an adaptation option in New
27 Zealand (Rouse et al., 2017; Hanna, 2019; Kool et al., 2020; Lawrence et al., 2020c) where protective
28 measures are transitional (DoC NZ, 2010) and where managed retreat has arisen from collaborative

Do Not Cite, Quote or Distribute                   11-59                            Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 governance (Owen et al., 2018). Remaining adaptation barriers are social or cultural (the absence of licence
 2 and legitimacy) and institutional (the absence of regulations, policies and processes that support changes to
 3 existing property rights and the funding of retreat) (O'Donnell and Gates, 2013; Tombs et al., 2018; Grace et
 4 al., 2019; O'Donnell et al., 2019) (high confidence).

 5

 6 Legacy development, competing public and private interests, trade-offs among development and
 7 conservation objectives, policy inconsistencies, short and long-term objectives, and the timing and scale of
 8 impacts, compound to create contestation over implementation of coastal adaptation (Mills et al., 2016b;
 9 McClure and Baker, 2018; Dedekorkut-Howes et al., 2020; McDonald, 2020; Schneider et al., 2020) (high
10 confidence). Legal barriers to coastal adaptation remain (Schumacher, 2020) with a risk that the courts
11 become decision makers (Iorns Magallanes et al., 2018) due to legislative fragmentation, status quo
12 leadership, lack of coordination between governance levels and agreement about who pays for what
13 adaptation (Waters et al., 2014; Boston and Lawrence, 2018; Palutikof et al., 2019a; Noy, 2020) (very high
14 confidence). The nexus of climate, law, place and property rights continues to expose people and assets to
15 ongoing sea-level rise (Johnston and France-Hudson, 2019; O'Donnell, 2019), especially where the risks of
16 sea-level rise are not being reflected in property valuations (Cradduck et al., 2020). Risk signalling through
17 land use planning, flooding events, and changes in insurance availability and costs, are projected to increase
18 recognition of coastal risks (Storey and Noy, 2017; CCATWG, 2018; Lawrence et al., 2018a; Harvey and
19 Clarke, 2019; Steffen et al., 2019; Cradduck et al., 2020; ICNZ, 2021) (medium confidence). Proactive local-
20 led engagement and strategy are key to effective adaptation and incentivising and supporting communities to
21 act(Gibbs, 2020; Schneider et al., 2020). Adopting `fit for purpose' decision tools that are flexible as sea
22 levels rise (11.7.3) can build adaptive capacity in communities and institutions (high confidence).

23

24 [END BOX 11.6 HERE]

25

26

27 11.3.6 Health and Wellbeing

28

29 11.3.6.1 Observed Impacts

30

31 There is ample evidence of health loss due to extreme weather in Australia and New Zealand, and rising
32 temperatures, changing rainfall patterns and increasing fire weather have been attributed to anthropogenic
33 climate change (11.2.1). Extreme heat leads to excess deaths and increased rates of many illnesses (Hales et
34 al., 2000; Nitschke et al., 2011; Lu et al., 2020). Between 1991 and 2011 it is estimated that 35-36% of heat-
35 related mortality in Brisbane, Sydney and Melbourne was attributable to climate change, amounting to about
36 106 deaths a year on average over the study period (Vicedo-Cabrera et al., 2021). Exposure to high
37 temperatures at work is common in Australia, and the health consequences may include more accidents,
38 acute heat stroke and chronic disease (Kjellstrom et al., 2016). Long-term rise in temperatures is changing
39 the balance of summer and winter mortality in Australia (Hanigan et al., 2021). The Black Summer wildfires
40 in Australia in 2019/2020 (Box 11.1) caused 33 deaths directly (Davey and Sarre, 2020) and exposed
41 millions of people to heavy particulate pollution (Vardoulakis et al., 2020). In the Australian States most
42 heavily affected by the fires, 417 deaths, 3151 hospital admissions for cardiovascular or respiratory
43 conditions, and about 1300 emergency department presentations for asthma are attributed to wildfire smoke
44 exposure (Borchers Arriagada et al., 2020). Immediate smoke-related health costs from the 2019-20 fires are
45 estimated at A$1.95 billion (Johnston et al., 2020).

46

47 Extreme heat is associated with decreased mental well-being, more marked in women than men (Ding et al.,
48 2016). Changing climatic patterns in Western Australia have undermined farmers' sense of identity and
49 place, heightened anxiety and increased self-perceived risks of depression and suicide (Ellis and Albrecht,
50 2017). Following the Black Saturday wildfires in Victoria in 2009, 10-15% of the population in the most
51 severely affected areas reported persistent fire-related post-traumatic stress disorder, depression and
52 psychological distress (Bryant et al., 2014). Repeated exposure to the threat of wildfires in Australia, either
53 directly (Box 11.1) or through media coverage (Looi et al., 2020) may compound effects on mental health. In
54 March 2017, 31,000 people in New South Wales and Queensland were displaced by Tropical Cyclone
55 Debbie. Six months post-cyclone, adverse mental health outcomes were more common among those whose
56 access to health and social care was disrupted (King et al., 2020).

57

Do Not Cite, Quote or Distribute  11-60       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Dengue fever remains a threat in northern Australia and variations in rainfall and temperature are related to
 2 disease outbreaks and patterns of spread, although most outbreaks are sparked by travellers bringing the
 3 virus into the country (Bannister-Tyrrell et al., 2013; Hall et al., 2021). Cases of dengue fever and other
 4 arboviral diseases have been increasing amongst recent arrivals to New Zealand from overseas, but to date
 5 there have been no reports of local transmission (Ammar et al., 2021).

 6

 7 In 2016 in New Zealand, it is estimated 6-8,000 people became ill due to contamination of the Havelock
 8 North water supply with the bacteria Campylobacter (Gilpin et al., 2020). The infection was traced to sheep
 9 faeces washed into the underground aquifer that feeds the town's (untreated) water supply after an
10 extraordinarily heavy rainfall event. This is not an isolated finding: increases in pediatric hospital admissions
11 are seen across New Zealand two days after heavy rainfall events (Lai et al., 2020).

12

13 11.3.6.2 Projected impacts

14

15 Climate change is projected to have detrimental effects on human health due to heat stress, changing rainfall
16 patterns including floods and drought, and climate-sensitive air pollution (including that caused by wildfires)
17 (high confidence). Vulnerability to detrimental effects of climate change will vary with socio-economic
18 conditions (high confidence).

19

20 The greatest number of people affected by compounding effects of heat, wildfires and poor air quality will be
21 in urban and peri-urban areas of Australia. By 2100 the proportion of all deaths attributable to heat in
22 Melbourne, Sydney and Brisbane may rise from about 0.5% to 0.8% (under RCP 2.6), or 3.2% (under RCP
23 8.5) (Gasparrini et al., 2017). Heat-wave related excess deaths in Melbourne, Sydney and Brisbane are
24 projected to increase to 300/year (RCP2.6) or 600/year (RCP8.5) during 2031-2080 relative to 142/year
25 during 1971-2020, assuming no adaptation and high population growth (Guo et al., 2018). High temperatures
26 amplify the risks due to local air pollution: without adaptation, ozone-related deaths in Sydney may increase
27 by 50-60 per year by 2070 (Physick et al., 2014).

28

29 Unless there is more effective control of nutrient run-off, bacterial contamination of drinking water supplies
30 is projected to increase due to more intense rainfall events, exacerbating risks to human health (Gilpin et al.,
31 2020, Lai, 2020 #2680), and higher temperatures will increase freshwater toxic blooms (Hamilton et al.,
32 2016).

33

34 Less certain climate change impacts include: surges in vector-borne diseases (medium confidence); threats to
35 mental health (medium confidence); reduction in winter mortality (medium confidence); emergence of new or
36 poorly understood weather-related threats (such as thunderstorm asthma or interactions between rising heat
37 and air pollution) (low confidence); and spill-over effects on health from global impacts of climate change
38 (e.g., on trade, conflict, migration) (low confidence).

39

40 In general, the area of Australia suitable for transmission of dengue is projected to increase (Zhang and
41 Beggs, 2018; Messina et al., 2019) but estimates of local disease risk vary considerably according to climate
42 change scenario and socio-economic pathways (Williams et al., 2016). The spread of Wolbachia amongst
43 Aedes mosquitoes in northern Australia has already reduced dengue transmission and may decrease the
44 influence of climate in the future (Ryan et al., 2019). In New Zealand, the risk of dengue remains low for the
45 remainder of this century (Messina et al., 2019). Higher temperatures and more intense rainfall may also
46 increase pollen production and the risk of allergic illness throughout the region (Haberle et al., 2014).

47

48 11.3.6.3 Adaptation

49

50 Strengthening basic public health services can rapidly reduce vulnerability to death and ill-health caused by
51 climate change, however this opportunity is often missed (very high confidence). The 2020 New Zealand
52 Health and Disability System Review pointed to short-comings in leadership and governance, structures that
53 embed health inequity, lack of transparency in planning and reporting, and under-investment in public health
54 personnel and systems (HDSR, 2020). An Australian study found that without deliberate planning the health
55 system `would only be able to deal with climate change in an expensive, ad hoc crisis management manner'
56 (Burton, 2014). In both Australia and New Zealand the COVID-19 epidemic has highlighted weaknesses in

Do Not Cite, Quote or Distribute  11-61       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 information systems, primary care for marginalized groups and inter-sectoral planning (Salvador-Carulla et
 2 al., 2020; Skegg and Hill, 2021): all these deficiencies are relevant to climate adaptation.

 3

 4 Underlying health and economic trends affect the vulnerability of the population to extreme weather (high
 5 confidence). Poor housing quality is a risk factor for climate-related health threats (Alam et al., 2016).
 6 Homeless people lack access to temperature-controlled or structurally safe housing, and often are
 7 excluded from disaster preparation and responses (Every, 2016). These inequalities are reversible. For
 8 example, a government partnership with social housing providers in Australia improved the thermal
 9 performance of housing for low-income tenants (Barnett et al., 2014a). A postcode-level analysis of the
10 vulnerability of urban populations to extreme heat in Australian capital cities (Loughnan et al., 2013) led to
11 the development of an interactive website for purposes of planning and emergency preparedness (Figure
12 11.5) as well as subsequent work on green urban design for cooler, more liveable cities (Tapper, In Press).

13

14

15

16 Figure 11.5: Housing and socio-economic disadvantage is correlated with the use of emergency services on hot days
17 (rho = 0.55, p<0.01). The spatial distribution of (A) a community vulnerability index (VI (PCA) by deciles and (B)
18 ambulance call-outs on days above daily mean of 34oC, in Brisbane, Australia. Ambulance call-out data are expressed
19 as deciles based on per-capita calls during 2003-2011 (Loughnan et al., 2013).

20

21

22 Heat-wave responses, from public education to formal heat-warning systems, are the best-developed element
23 of adaptation planning for health in Australia, but many metropolitan centres are still not covered (Nicholls
24 et al., 2016; Nitschke et al., 2016) (high confidence). Air conditioning (AC) in Australian homes reduces
25 mortality in heat-waves by up to 80% (Broome and Smith, 2012) but heavy reliance on AC carries risks. It is
26 estimated that a power outage on the third day of extreme heat-waves would result in an additional 10­21
27 deaths in Adelaide, 24­47 in Melbourne and 7­13 in Brisbane (Nairn and Williams, 2019). Multiple
28 interventions at the landscape, building and individual scale are available to reduce the negative health
29 effects of extreme heat (Jay et al., 2021)

30

31 Heat extremes receive most policy attention, but the numbers of deaths are less than those resulting from
32 more frequent exposures to moderately high temperatures (Longden, 2019). Melbourne provides a case study
33 in long-term planning for cooler cities, with its Urban Forest Strategy (Gulsrud et al., 2018). Australian

Do Not Cite, Quote or Distribute  11-62       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 workers' perceptions of heat and responses to high temperatures show that heat policies on their own are
 2 insufficient for full protection; workers also require knowledge and agency to slow down or take breaks on
 3 their own initiative (Singh et al., 2015; Lao et al., 2016).

 4

 5 The first national climate change risk assessment in New Zealand (MfE, 2020a) highlighted the risk to
 6 potable water supplies. An inquiry into the Havelock North outbreak recommended that all registered
 7 drinking water supplies (which supply about 80% of the national population) in New Zealand should be
 8 disinfected and have stronger oversight by a national regulatory body (Government Inquiry into Havelock
 9 North Drinking Water, 2017). The use of local and Indigenous knowledge strengthens interventions to
10 protect water supplies to remote settlements that may be affected by climatic changes (Henwood, 2019).

11

12 Adaptation requires better protection of health facilities and supply chains, but hospital managers seldom
13 have capacity to invest in long-term improvements in infrastructure (Loosemore et al., 2014). However,
14 health services in the region are required to prepare disaster plans: these could be expanded to explicitly
15 cover health adaptation and local threats from climate change, including flooding events (Rychetnik et al.,
16 2019).

17

18 11.3.7 Tourism

19

20 11.3.7.1 Observed Impacts

21

22 Tourism is a major economic driver in the region, accounting for 3% (Australia) and 6% (New Zealand) of
23 GDP pre-COVID-19 (WTTC, 2018). Climate change is having significant impacts on tourism due to the
24 heavy reliance of the sector on natural heritage and outdoor attractions (11.3.1; Box 11.2). Furthermore, as
25 Australia and New Zealand are both long-haul destinations, a global increase in `flygskam' (flight shame)
26 will to impact travel patterns (Becken et al., 2021).

27

28 Impacts of climate change are being observed across the tourism system (Scott et al., 2019a) (high
29 confidence), most notably the Great Barrier Reef (Box 11.2) (Ma and Kirilenko, 2019). Australia's ski
30 industry is very sensitive to climatic change, due to reduction in snow depth and the length of the snow
31 season (Table 11.2) (Steiger et al., 2019; Knowles and Scott, 2020). The 2019-2020 summer wildfires (Box
32 11.1), impacted tourism and travel infrastructure, affecting air quality, vineyards and wineries (CoA, 2020e;
33 Filkov et al., 2020). Global media coverage of the wildfires, alongside Australia's climate change policy
34 response, profoundly and negatively, affected Australia's destination image (Schweinsberg et al., 2020; Wen
35 et al., 2020). In New Zealand's South Island, Fox and Franz Josef Glaciers have retreated approximately
36 700m since 2008, with ice melt and retreat resulting in increased rock fall risks and negatively affecting the
37 tourist experience (Purdie, 2013; Stewart et al., 2016; Wang and Zhou, 2019). The West Coast of New
38 Zealand is extremely prone to flooding events impacting amenity values and access (Paulik et al., 2019b).
39 Damage to tracks, huts and bridges have closed popular destinations, including the Hooker Glacier and the
40 popular Routeburn and Heaphy Tracks during heavy rainfall events (Christie et al., 2020). Climate-driven
41 damage is motivating `last chance' tourism to see key natural heritage and outdoor attractions, e.g. Great
42 Barrier Reef (Piggott-McKellar and McNamara, 2016) and Franz and Fox Glaciers (Stewart et al., 2016).

43

44 11.3.7.2 Projected Impacts

45

46 Widespread impacts from projected climate change are very likely across the tourism sector. The World
47 Heritage listed Kakadu National Park in Australia is projected to experience increasing severity of cyclones
48 (Turton, 2014) and sea-level rise is projected to affect freshwater wetlands (11.3.1.2; Table 11.5) (McInnes
49 et al., 2015) and Indigenous rock art (Higham et al., 2016; Hughes et al., 2018a). The projected increase in
50 the number of hot days in northern and inland Australia may impact the attractiveness of the region for
51 tourists (Amelung and Nicholls, 2014; Webb and Hennessy, 2015). Coastal erosion and flooding of
52 Australasian beaches due to sea-level rise and intensifying storm activity is estimated to increase by 60% on
53 the Sunshine Coast by 2030 causing significant damage to tourist-related infrastructure (Hughes et al.,
54 2018a). Urgent `hard' and `soft' adaptation strategies are projected to help reduce sea-level rise impacts
55 (Becken and Wilson, 2016).

56

Do Not Cite, Quote or Distribute  11-63       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Glacier tourism, a multimillion-dollar industry in New Zealand, is potentially under threat because glacier
 2 volumes are projected to decrease (Purdie, 2013) (very high confidence). Glacier volume reductions of 50­
 3 92% by 2099 relative to present reflect the large range of temperature projections between RCP2.6 and
 4 RCP8.5. Under RCP2.6 at 2099, the glaciers retain a similar configuration to present, although clean-ice
 5 glaciers will retreat significantly. For RCP4.5, RCP6.0 and RCP8.5, the clean-ice glaciers will retreat to
 6 become small remnants in the high mountains (Anderson et al. 2021).

 7

 8 Snow skiing faces significant challenges from climate change (high confidence). In Australia, the annual
 9 maximum snow depth is estimated to decrease from current levels by 15% (2030) and 60% by 2070 (SRES
10 A2) (Di Luca et al., 2018). By 2070-2099, relative to 2000-2010, the length of the Victorian ski-season is
11 projected to contract by 65-90% under RCP8.5 (Harris et al., 2016). The New Zealand tourism destination of
12 Queenstown is expected to experience declining snowfall, increased wind and more severe weather events
13 (Becken and Wilson, 2016). Ski tourism stakeholders have been responding to longer-term climate risks with
14 an increase in snow-making machines in New Zealand since 2013 (Hopkins, 2015) and in Australia (Harris
15 et al., 2016).

16

17 11.3.7.3 Adaptation

18

19 Current snow-making technologies are expected to sustain the ski industry until mid-century. However, with
20 warmer winter temperatures and declining water availability, snow-making is projected to decrease to half at
21 most resorts by 2030 (Harris et al., 2016). New Zealand's ski industry may benefit from Australian skiers
22 visiting New Zealand, due to lower relative vulnerability (Hopkins, 2015). However, tourists may substitute
23 destinations or ski less in the absence of snow (medium agreement, limited evidence) (Cocolas et al., 2015;
24 Walters and Ruhanen, 2015).

25

26 With the exception of the ski industry (Becken, 2013; Hopkins, 2015), tourism stakeholders generally focus
27 on coping with short-term weather events, rather than longer-term climate risks, but do exhibit high adaptive
28 capacity by diversifying their activities (Stewart et al., 2016). Post Covid-19 pandemic economics and
29 recovery policies challenge this sector's prospects, and the combination of COVID-19 and climate change
30 (e.g. fires, floods) has also highlighted the need for the tourism sector to be able to respond to multiple,
31 overlapping crises.

32

33 There is limited evidence that research into the impact of climate change on tourism in Australia and New
34 Zealand is translating into policy or action (Moyle et al., 2017). New Zealand government tourism sector
35 strategies acknowledge this and the need for greater understanding of climate change for the sector, (TIA,
36 2019), but do not offer solutions (MBIE, 2019b; MfE, 2020a). The COVID-19 pandemic and the global
37 pause of international travel offers an opportunity to potentially `reset' tourism to account for the impacts of
38 climate change (Prideaux et al., 2020).

39

40 11.3.8 Finance

41

42 11.3.8.1 Observed Impacts

43

44 The finance sector has significant exposure to climate variability and extreme events (high confidence).
45 Aggregated insured losses from weather-related hazard events from 2013-2020 were almost A$15 billion for
46 Australia (1.2% of GDP) and almost NZ$1 billion for New Zealand (0.4% of GDP) (ICA, 2020a; NIWA,
47 2020). However, there is no trend in normalised losses because the rising insurance costs are being driven by
48 more people living in vulnerable locations with more to lose (McAneney et al., 2019). In New Zealand, two
49 major hailstorms during 2014-2020 and three major floods during 2019-2021 caused significant insurance
50 losses (ICNZ, 2021). Insured losses exceeded NZ$472 million for the 12 costliest floods from 2007-2017, of
51 which NZ$140 million could be attributed to anthropogenic climate change (Frame et al., 2020). In
52 Australia, insured damage was almost A$1.0 billion for the Queensland hailstorm in 2020, A$1.7 billion for
53 east coast flooding in 2020, A$2.3 billion for the 2019-2020 fires, A$2.3 billion for the Queensland
54 hailstorm in 2019, A$1.2 billion for the north Queensland floods in 2019, A$1.4 billion for the NSW
55 hailstorm in 2018, A$1.8 billion for Cyclone Debbie in 2017 and A$1.5 billion for the Brisbane hailstorm in
56 2014 (ICA, 2020b). The insured loss from the seven costliest hailstorms in Australia from 2014-2021
57 totalled A$7.6 billion (ICA, 2021).

Do Not Cite, Quote or Distribute  11-64       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1

 2 Some homes in the highest risk areas tend to be in lower socio-economic groups that may not buy insurance
 3 (Actuaries Institute, 2020). For example, one quarter of residents that experienced loss or damage in the
 4 2019 Townsville floods did not have insurance (ACCC, 2020). Under-insurance reduces people's capacity to
 5 recover from adverse events, while over-reliance on private insurance undermines collective disaster
 6 recovery efforts (Lucas and Booth, 2020). In Australia, those in high-risk areas minimise house and contents
 7 insurance for financial reasons (Booth and Harwood, 2016; Osbaldison et al., 2019; Actuaries Institute,
 8 2020). Insurance premiums in northern Australia are almost double those in the rest of Australia, and rising,
 9 mainly due to cyclone damage (ACCC, 2020).

10

11 11.3.8.2 Projected Impacts

12

13 Risks for the finance sector are projected to increase (medium confidence). The potential impact of increased
14 coastal and inland flooding, soil desiccation and contraction, fire and wind could lead to higher insurance
15 costs, reduced property values and difficulty for some customers to service loans (CBA, 2018). Under a high
16 emission scenario (RCP8.5), estimated annual losses to home-lending customers may increase 27% by 2060,
17 and the proportion of properties with high credit risk may rise from 0.01% in 2020 to 1% in 2060, assuming
18 no change in the portfolio (CBA, 2018). In New Zealand, weather-related insurance claims between 2000­
19 2017 totaled NZ$450 million, 40% of which were due to extreme rainfall. Using six climate model
20 projections of extreme rainfall, the insured damage is projected to increase by 7% (RCP2.6) to 8% (RCP8.5)
21 by 2020­2040 and 9% (RCP2.6) to 25% (RCP8.5) by 2080­2100, relative to 2000­2017 (Pastor-Paz et al.,
22 2020). By 2050­2070, tropical cyclone risk for properties not in flood plains or storm surge zones in south-
23 east Queensland may increase by 33% under a 2°C scenario, and by 317% under a 3°C scenario for properties
24 in flood plains and storm surge zones (IAG, 2019).

25

26 11.3.8.3 Adaptation

27

28 Banks, insurers and investors increasingly recognise the risks posed by climate change to their businesses
29 (Paddam and Wong, 2017) (high confidence). Collaborations between banks, insurers and superannuation
30 funds in Australia and New Zealand are driving efforts aimed at achieving the Paris Agreement goals,
31 including the New Zealand Centre for Sustainable Finance and Australian Sustainable Finance Initiative
32 (AFSI, 2020; TAO, 2020; NZCFSF, 2021). Company directors including superannuation fund directors have
33 legal obligations to disclose and appropriately manage material financial risks (Barker et al., 2016; Hutley
34 and Davis, 2019). Financial regulators are aware of climate risks for financial stability and financial
35 institutions (RBNZ, 2018; RBA, 2019) and are closely supervising climate risk disclosure practices (TCFD,
36 2017; RBNZ, 2018; APRA, 2019; CMSI, 2020; IGCC, 2021b). In Australia, regulatory action (APRA, 2021)
37 includes issuing prudential guidelines for financial institutions on managing climate risk, aligned with
38 guidelines developed by the Climate Measurement Standards Initiative (NESP ESCC, 2020). In New
39 Zealand, the Financial Sector (Climate-related Disclosure and Other Matters) Amendment Bill aims to
40 ensure that the effects of climate change are routinely considered in business, investment, lending, and
41 insurance underwriting decisions
42 (NZ Government, 2021).

43

44 Banks and insurers are beginning to undertake climate risk analyses (CRO Forum, 2019; Bruyère et al.,
45 2020) and disclose their risks (Paddam and Wong, 2017; ANZ, 2018; CBA, 2018). For example, the
46 agricultural banking sector has analysed climate risk and embedded climate adaptation financing into its risk
47 scoring and lending practices (CBA, 2019). However, the overall number of disclosures continues to lag
48 expectations, suggesting the need for mandatory climate risk disclosure in Australia (IGCC, 2021a).

49

50 Climate adaptation finance is not evident (medium confidence). There is an adaptation finance gap (Mortimer
51 et al. 2020). Private sector initiatives are beginning to emerge through large scale projects or public-private
52 partnerships, such as the Queensland Betterment Fund (Banhalmi-Zakar et al., 2016; Ware and Banhalmi-
53 Zakar, 2020). Addressing investor pressure (IGCC, 2017) could increase investment in adaptation. However,
54 ongoing policy uncertainty in Australia continues to be the key barrier to allocating further capital to invest
55 in climate solutions for 70% of investors (IGCC, 2021a).

56

Do Not Cite, Quote or Distribute  11-65       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 Current and future insurance affordability pressures could be addressed by increased mitigation, revisions to
 2 building codes and standards, and better land-use planning (ACCC, 2020; Actuaries Institute, 2020). In New
 3 Zealand, insurance signals are motivating the government to address adaptation funding mechanisms
 4 (Boston and Lawrence, 2018; CCATWG, 2018). Some insurers offer premium discounts to customers with
 5 reduced risk (Drill et al., 2016) with increasing premiums reflecting known risk and no cover for some
 6 hazards in risky locations (CCATWG, 2017). Special excess payments are available for flood hazard so
 7 customers take responsibility for part of the claim, with increasing premiums to reflect known and
 8 foreseeable risk, and downgrading cover from replacement value to market value (Bruyère et al., 2020).
 9 Retreat by private insurers from risky locations could increase the unfunded fiscal risk to the government
10 (Storey and Noy, 2017) creating moral hazard (Boston and Lawrence, 2018). The litigation risk from failing
11 to take adaptation action (Hodder, 2019) could affect financial markets and government policy settings,
12 creating cascading impacts across society (Lawrence et al., 2020b)(CRO Forum, 2019). For some climate
13 risks, national governments act as "last resort" insurers (CCATWG, 2017), but this could become
14 unsustainable (CRO Forum, 2019).

15

16 11.3.9 Mining

17

18 Many mines are exposed and sensitive to climate extremes (high confidence), but there is little available
19 research on climate change impacts (Odell et al., 2018). Most Australian mines face higher temperatures,
20 cyclones, erosion and landslides, and hazards such as sea-level rise and storms across their supply chains,
21 including ports (Cahoon et al., 2016). Impacts include operational disruptions such as acute drainage
22 problems (Loechel and Hodgkinson, 2014) and heat-induced illness, irritation and absenteeism among
23 workers (McTernan et al., 2016), lost revenue and increased costs (Pizarro et al., 2017).

24

25 Damage and disruption from climate impacts can cost operators billions of dollars (Cahoon et al., 2016).
26 Climatic extremes increase the risk and impact of spillages along transportation routes (Grech et al., 2016)
27 exacerbate mining's effects on hydrology, ecosystems, and air quality (Phillips, 2016; Ali et al., 2018);
28 increase contamination risks (Metcalfe and Bui, 2016); and disrupt and slow mine site rehabilitation
29 (Wardell-Johnson et al., 2015; Hancock et al., 2017). Adaptations such as improved water management are
30 emerging slowly (Gasbarro et al., 2016; Becker et al., 2018). Some companies are spatially diversifying and
31 relocating (Hodgkinson et al., 2014). Others are replacing workers with automation and remote operations
32 (Halteh et al., 2018; Keenan et al., 2019).

33

34 11.3.10 Energy

35

36 Australia's energy generation is a mix of coal (56%), gas (23%) and renewables (21%) (DISER, 2020), with
37 ageing coal-fired infrastructure being replaced with a growing proportion of renewable and distributed
38 energy resources (AEMO, 2018). In New Zealand, 60% of energy generation comes from hydro-electricity
39 and 15% from geothermal (MBIE, 2021), with coal (2%) and gas (13%) generation capacity to be retired,
40 and total renewable energy to increase from 82% in 2017 to around 95% by 2050, mostly through wind
41 generation (MBIE, 2019a).

42

43 11.3.10.1 Observed Impacts

44

45 The energy sector is highly vulnerable to climate change (high confidence). Oil and gas systems are
46 vulnerable to storms, fires, drought, floods, sea-level rise, extreme heat and fires which can damage
47 infrastructure, slow production, and add to operational costs (Smith, 2013). The electricity system is
48 vulnerable to high temperatures reducing generator and network capacity and increasing failure rates and
49 maintenance costs(AEMO, 2020a). Fires (including those sparked by electrical distribution lines) pose risks
50 to assets, smoke can cause electricity transmission to trip, high winds reduce wind-energy capacity and
51 threaten the integrity of transmission lines, low rainfall reduces hydro-energy capacity and increases the
52 demand for desalination energy, higher sea-level may affect some low-lying generation, distribution and
53 transmission assets, and compound extreme weather events can cause outages (Vose and Applequist, 2014;
54 Lawrence et al., 2016; AEMO, 2020b; AEMO, 2020a; ESCI, 2021). For example, in September 2016, a
55 major windstorm in South Australia damaged 23 transmission towers and cut power to over 900,000
56 households. In February 2017, the South Australian energy system failed to cope with a heatwave-related
57 jump in demand, causing power cuts to 40,000 homes (Steffen et al., 2017). In April 2018, a storm over

Do Not Cite, Quote or Distribute  11-66       Total pages: 151
FINAL DRAFT                       Chapter 11                                  IPCC WGII Sixth Assessment Report

1 Auckland New Zealand left 182,000 properties without power (Bell, 2018). The 2019/20 Australian
2 heatwaves and fires caused widespread blackouts that disrupted communications, transport, and emergency
3 response capacity (Box 11.1).

4

5 11.3.10.2 Projected Impacts

 6

 7 Risks for the energy sector are projected to increase with climate change (medium confidence). Projected
 8 increases in the frequency and intensity of heatwaves, fires, droughts and wind-storms would increase risks
 9 for energy supply and demand (AEMO, 2020b; ESCI, 2021). Households are unevenly vulnerable to energy
10 sector risks due to varying housing quality and health dependencies (11.3.6). In New Zealand, a warmer
11 climate and increasing energy efficiency is projected to marginally reduce annual average peak electricity
12 heating demand (Stroombergen et al., 2006; MBIE, 2019a). Winter and spring inflows to main hydro lakes
13 are projected to increase 5-10% and may reduce hydroelectric energy vulnerability (McKerchar and Mullan,
14 2004; Poyck et al., 2011; Stevenson et al., 2018). However, major electricity supply disruptions are projected
15 to increase as dependence on electricity grows from 25% of total energy in 2016 to 58% in 2050
16 (Transpower, 2020).

17

18 In Australia, the total heating and cooling energy demand of 5-star energy-rated houses is projected to
19 change by 2100 (Wang et al., 2010). At 2°C global warming, the estimated change in demand is ­27% in
20 Hobart, ­21% in Melbourne, +61% in Darwin, +67% in Alice Springs and +112% in Sydney. For a 4°C
21 global warming, the changes are ­48%, ­14%, +135%, +213% and +350% respectively.

22

23 11.3.10.3 Adaptation

24

25 Options to manage risks include adaptation of energy markets, integrated planning, improved asset design
26 standards, smart-grid technologies, energy generation diversification, distributed generation (e.g. roof-top
27 solar, micro-grids), energy efficiency, demand management, pumped hydro storage, battery storage, and
28 improved capacity to respond to supply deficits and balance variable energy resources across the network
29 (Table 11.8) (high confidence). With increasing electrification, diversification and resilience can contribute
30 to security of supply as fossil fuels are retired from the energy mix (AEMO, 2020b). In Australia, the AEMO
31 (2020) Integrated System Plan has evaluated various options, costs and benefits. Risks associated with an
32 increasing reliance on weather-dependent renewable energy (e.g. solar, wind, hydro) (ESCI, 2021) can be
33 managed through strong long-distance interconnection via high voltage powerlines and storage (Blakers et
34 al., 2017; Blakers et al., 2021; Lu et al., 2021). However, implementation of adaptation options remains
35 inadequate (Gasbarro et al., 2016).

36

37

38 Table 11.8: Adaptation options for the energy sector.

Adaptation options                                                            References

Diversification of electricity supplies geographically and technically,       (AEMO, 2020b)
including distributed energy resources and variable renewable energy

Integrated planning, improved asset design and management, and disaster       (AEMO, 2020b; Transpower, 2020)
recovery to build resilience to more extreme weather

Augmentation of transmission grid to support change in generation mix         (Blakers et al., 2017; ICCC, 2019;
using interconnectors and renewable energy zones, coupled with energy         AEMO, 2020b)
storage, adds capacity and helps balance variable resources across the
network                                                                       (Bridge et al., 2018; AEMO, 2020b)

Climate change risks included in the design, location, and rating of future
infrastructure and consideration of the implications for future transmission
developments

Increased design and construction standards, flood defence measures,          (Smith, 2013; Gasbarro et al., 2016)
insurance, improved water efficiency, improved insulation of super-cooled
LNG processes, more efficient air conditioning and creating fire breaks for
the oil and gas sector

Do Not Cite, Quote or Distribute  11-67                                                      Total pages: 151
   FINAL DRAFT                       Chapter 11                                IPCC WGII Sixth Assessment Report

   Technological developments to strengthen existing resilience under climate  (Evans et al., 2018)
   change that reinforces the relative advantage of Western Australia and
   Tasmania for new wind energy installations

   Energy generation diversity, demand management, pumped hydro storage        (Keck et al., 2019; Transpower, 2020)
   and battery storage

   Tools and strategies to manage winter energy deficits and dry years         (Transpower, 2020)
   alongside renewable electricity generation deployment

   Improved insulation and heating of buildings, and flexible electricity      (Stroombergen et al., 2006; MBIE,
   consumption to reduce the significance of winter electricity demand peak    2019a; Transpower, 2020)

1

2

3 11.3.11 Detection and Attribution of Observed Climate Impacts

 4

 5 Detection and attribution of observed climate trends and events is called `climate attribution'. This has been
 6 assessed by IPCC Working Group I (Gutiérrez et al., 2021; Ranasinghe et al., 2021; Seneviratne et al., 2021)
 7 and summarised in IPCC Working Group 2 Chapter 16. Trends that have been formally attributed in part to
 8 anthropogenic climate change include regional warming trends and sea-level rise, decreasing rainfall and
 9 increasing fire risk in southern Australia. Events include extreme rainfall in New Zealand during 2007-2017,
10 the 2007/8 and 2012/13 droughts in New Zealand, high temperatures in Australia during 2013-2020, the
11 2016 northern Australian marine heatwave, the 2016/2017 and 2017/18 Tasman Sea marine heatwaves, and
12 2019/2020 fires in Australia.

13

14 Detection and attribution of climate impacts on natural and human systems is called `impact attribution'.
15 This often involves a two-step approach (joint attribution) that links climate attribution to observed impacts.
16 Impact attribution is complicated by confounding factors, e.g. changes in exposure arising from population
17 growth, urban development and underlying vulnerabilities.

18

19 Impact attribution has been considered in Sections 11.3.1 to 11.3.10 and summarised in Table 11.9. More
20 literature is available for natural systems than human systems, which represents a knowledge gap rather than
21 an absence of impacts that are attributable to anthropogenic climate change. Fundamental shifts in the
22 structure and composition of some ecosystems are partly due to anthropogenic climate change (high
23 confidence). In human systems, the costs of droughts and floods in New Zealand, and heat-related mortality
24 and fire damage in Australia, are partly attributed to anthropogenic climate change (medium confidence).

25

26

27 Table 11.9: Examples of observed impacts that can be partly attributed to climate change.

   Impact                                                                                       Source

   Mass bleaching of the Great Barrier Reef in 2016/2017 due to a marine heatwave               Box 11.2

   In the New Zealand Southern Alps, extreme glacier mass loss was at least six times more      11.2.1, 11.3.3
   likely in 2011, and ten times more likely in 2018, due to warming                            Table 11.4

   In the Australian Alps bioregion, loss of habitat for endemic and obligate species due to
   snow loss and increases in fire, drought and temperature

   In the Australian wet tropics world heritage area, some vertebrate species have declined in  Table 11.4
   distribution area and population size due to increasing temperatures and length of dry
   season

   Extinction of Bramble Cay melomys due to loss of habitat caused by storm surges and sea- Table 11.4
   level rise in Torres Strait

   In New Zealand, increasing invasive predation pressure on endemic forest birds surviving Table 11.4
   in cool forest refugia due to anthropogenic warming

   In New Zealand, erosion of coastal habitats due to more severe storms and sea-level rise Table 11.4, Box 11.6

   Do Not Cite, Quote or Distribute  11-68                                                           Total pages: 151
FINAL DRAFT                       Chapter 11                       IPCC WGII Sixth Assessment Report

In Australia, estuaries warming and freshening with decreasing pH                            Table 11.6

Changes in life-history traits, behaviour or recruitment of fish and invertebrates due to    Table 11.6
ocean acidification or warming, severe decline in recruitment of coral on the Great Barrier
Reef due to ocean warming, aquaculture stock deaths due to heat stress

New diseases and toxins due to warming and extension of East Australian Current              Table 11.6
Changes in almost 200 marine species distributions and abundance due to ocean warming        Table 11.6

Temperate marine species replaced by seaweeds, invertebrates, corals and fishes              Table 11.6
characteristic of subtropical and tropical waters

River flow decline in southern Australia is largely due to the decline in cool season rainfall 11.3.3
partly attributed to anthropogenic climate change

In New Zealand, the 2007/08 drought and the 2012/13 drought were 20% attributed to           11.3.3
anthropogenic climate change

In New Zealand, about 30% of the insured damage for the 12 costliest flood events from       11.3.8
2007-2017 can be attributed to anthropogenic climate change

          In Australia, 35-36% of heat-related excess mortality in Melbourne, Sydney and Brisbane 11.3.6
          from 1991-2018 can be attributed to anthropogenic climate change

 1

 2

 3 11.4 Indigenous Peoples

 4

 5 Indigenous perspectives of well-being embrace physical, social, emotional and cultural domains,
 6 collectiveness and reciprocity, and more fundamentally connections between all elements across the past,
 7 present and future generations (Australia. NAHS Working Party, 1989; MfE, 2020a). Changing climate
 8 conditions are expected to exacerbate many of the social, economic and health inequalities faced by
 9 Aboriginal and Torres Strait Islander Peoples in Australia and Mori in New Zealand (Bennett et al., 2014;
10 Hopkins et al., 2015; AIHW, 2016; Lyons et al., 2019) (high confidence). As a consequence, effective policy
11 responses are those that take advantage of the interlinkages and dependencies between mitigation, adaptation
12 and Indigenous Peoples' wellbeing (Jones, 2019) and those that address the transformative change needed
13 from colonial legacies (Hill et al., 2020) (high confidence). There is a central role for Indigenous Peoples in
14 climate change decision making that helps address the enduring legacy of colonisation through building
15 opportunities based on Indigenous governance regimes, cultural practices to care for land and water, and
16 intergenerational perspectives (Nursey-Bray et al., 2019; Petzold et al., 2020) (Cross-Chapter Box INDIG in
17 Chapter 18) (very high confidence).

18

19 11.4.1 Aboriginal and Torres Strait Islander Peoples of Australia

20

21 The highly diverse Aboriginal and Torres Strait Islander Peoples of Australia have survived and adapted to
22 climate changes such as sea-level rise and extreme rainfall variability during the late Pleistocene era, through
23 intimate place-based Indigenous Knowledge in practice and while losing traditional land and sea Country
24 ownership (Liedloff et al., 2013) (Cross-Chapter-Box INDIG in Chapter 18) including during the Late
25 Pleistocene era (Golding and Campbell, 2009; Nunn and Reid, 2016). They belong to the world's oldest
26 living cultures, continually resident in their own ancestral lands, or `country', for over 65,000 years
27 (Kingsley et al., 2013; Marmion et al., 2014; Nagle et al., 2017; Tobler et al., 2017; Nursey-Bray and
28 Palmer, 2018). The majority of the Australian Indigenous Peoples live in urban areas in southern and eastern
29 Australia, but are the predominant population in remote areas.

30

31 Climate-related impacts on Aboriginal and Torres Strait Islander Peoples, Countries (traditional estates) and
32 cultures have been observed across Australia and are pervasive, complex and compounding(Green et al.,
33 2009) (11.5.1) (high confidence). For example, loss of bio-cultural diversity, nutritional changes through
34 availability of traditional foods and forced diet change, water security, and loss of land and cultural resources
35 through erosion and sea-level rise (Table 11.10) ](TSRA, 2018). Moreover, these impacts are being
36 experienced now particularly in low-lying geographical areas- especially in the Torres Strait Islands (Mosby,

Do Not Cite, Quote or Distribute  11-69                                                                Total pages: 151
FINAL DRAFT                            Chapter 11  IPCC WGII Sixth Assessment Report

1 2012; Kelly, 2014; Murphy, 2019; Hall et al., 2021). Estimates of the loss from fire impacts on ecosystem
2 services that contribute to the wellbeing of remotely-located Indigenous Australians were found to be higher
3 than the financial impacts from the same fires on pastoral and conservation lands (Sangha et al., 2020) and
4 could increase with both financial and non-financial impacts (Box 11.1).

5

6

7 Table 11.10: Climate-related impacts on Aboriginal and Torres Strait Islander Peoples, country and cultures.

Impacts                 Implications

Loss of bio-cultural    Healthy country is critical to Indigenous Australians' livelihoods, caring for country
diversity (land, water  responsibilities, health and wellbeing. Damage to land can magnify the loss of spiritual
and sky) (medium        connection to land from dispossession from traditional Country and leads to disruption of
confidence)             cultural structures. Climate change impacts can exacerbate and/or accelerate existing threats
                        of habitat degradation and biodiversity loss, and create challenges for traditional stewardship
                        of landscapes (Mackey and Claudie, 2015)

Climate-driven loss     Traditional coastal lands lost through erosion and rising sea level, with associated mental
of native title and     health implications from loss of cultural and traditional artefacts and landscapes, including the
other customary         destruction and exhumation of ancestral graves and burial grounds. This is also occurring and
lands (medium           predicted to intensify in the low-lying islands of the Torres Strait (TSRA, 2018; Hall et al.,
confidence)             2021) and was also noted during the extreme bushfires in Eastern Australia in late 2019 and
                        early 2020.

Changing availability Human health impacts can be exacerbated by climate change through changing availability of

of traditional foods traditional foods and medicines, while outages and high costs of electricity can limit storage

and forced diet         of fresh food and medication (Kingsley et al., 2013; Spurway and Soldatic, 2016; Hall and

change (medium          Crosby, 2020)

confidence)

Changing climatic       Climate change-induced sea-level rise and saltwater intrusion can limit the capacity for
conditions for          traditional Indigenous floodplain pastoralism, and also affect food security, access and
subsistence food        affordability to healthy, nutritional food (Ligtermoet, 2016; Spurway and Soldatic, 2016)
harvesting (medium
confidence)             Increasing frequency or intensity of extreme weather events (floods, droughts, cyclones,
                        heatwaves) can cause disaster responses in remote communities, including infrastructure
Extreme weather         damage of essential water and energy systems and health facilities (TSRA, 2018; Hall and
events triggering       Crosby, 2020)
disasters (high
confidence)

Heatwave impacts on     Heatwaves can occur in many regions. Tropical regions can experience prolonged seasons of
human health (high      high temperatures and humidity levels, resulting in extreme heat stress risks. For example, the
confidence)             Torres Strait Island are already categorised under the U.S. National Oceanic and Atmospheric
                        Administration (NOAA) Heat Index as a danger zone for extreme human health risk during
                        Summer (TSRA, 2018)

Health impacts from     Climate change can change exposure and increase risk for remote Indigenous Peoples to
changing conditions     infection from waterborne and insect-borne diseases, especially if medical services are limited
for vector-borne        or damaged by extreme weather events. For example, in the Torres Strait Islands the changing
diseases (high          climate is affecting the range and extension of the Aedes albopictus and Aedes aegypti
confidence)             mosquitoes that can carry and transmit dengue and other viruses (Horwood et al., 2018;
                        TSRA, 2018)

Unadaptable             Poorly-designed, inferior quality and unmaintained housing can create health challenges for
infrastructure for      tenants in extreme heat (Race et al., 2016). Essential community-scale water and energy
changing                service infrastructure, unpaved roads, sea walls and storm water drains can fail in extreme
environmental           weather events (McNamara et al., 2017)
conditions (high
confidence)

Do Not Cite, Quote or Distribute       11-70       Total pages: 151
    FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

    Drinking water    Predicted continued increases in arid conditions in Australia are expected to reduce the
    security (medium  recharge rate of finite groundwater supplies (Barron et al., 2011). For remote communities
    confidence)       reliant on groundwater for drinking supplies, this water insecurity creates vulnerabilities from
                      over-extraction and lack of access (Jackson et al., 2019; Hall and Crosby, 2020). This
                      groundwater can also have microbial contamination from sewage and chemicals supporting
                      bacterial growth, such as high iron levels supporting the growth of Burkholderia pseudomallei
                      that causes melioidosis in humans and animals (Kaestli et al., 2019). In the Torres Strait,
                      increasing reliance on desalination for drinking water raises costs for fuel and its associated
                      transport (Beal et al., 2018)

1

 2

 3 Due to ongoing impacts of colonisation, Aboriginal and Torres Strait Islander Peoples have, on average,
 4 lower income, poorer nutrition, lower school outcomes and employment opportunities, and higher
 5 incarceration and removal of children than non-Indigenous Australians, represented in high comorbidities of
 6 chronic diseases and mental health impacts (Marmot, 2011; Green and Minchin, 2014; AIHW, 2015). This
 7 relative poverty can reduce climate-adaptive capacities while exacerbating climate change vulnerabilities
 8 (Nursey-Bray and Palmer, 2018). In remote Country, this can combine with lack of security for food and
 9 water, non-resilient housing and extreme weather events, contributing to migration off traditional Country
10 and into towns and cities- with flow-on social impacts such as homelessness, dislocation from community
11 and family, and disconnection from country and spirituality (Mosby, 2012; Brand et al., 2016).

12

13 Recognition of the role Aboriginal and Torres Strait Islander Peoples in identifying solutions to the impacts
14 of climate change is slowly emerging (UN, 2018) having been largely excluded from meaningful
15 representation from the conception of climate change dialogue, through to debate and decision-making
16 (Nursey-Bray et al., 2019). Honouring the United Nations' Declaration on the Rights of Indigenous Peoples
17 and social justice values would support self-determination and the associated opportunity for Indigenous
18 Australians to develop adaptation responses to climate change (Langton et al., 2012; Nursey-Bray and
19 Palmer, 2018; Nursey-Bray et al., 2019), including the adaptive capacity opportunities available through
20 Indigenous Knowledge (Liedloff et al., 2013; Petheram et al., 2015; Stewart et al., 2019) (Cross-Chapter Box
21 INDIG in Chapter 18). The Uluru Statement from the Heart proposes a pathway and roadmap forward for
22 enhanced representation of Aboriginal and Torres Strait Islander Peoples in decision-making in Australia
23 (Ululru Statement, 2017). Table 11.11 provides examples of traditional Indigenous practices of adaptation to
24 a changing climate. However, due to Indigenous methods of knowledge sharing and knowledge holding,
25 such knowledge relies disproportionately on elders and seniors, who form a very small portion of the total
26 Aboriginal and Torres Strait Islander Peoples of Australia, and is limited in the formal literature (ABS,
27 2016).

28

29

30 Table 11.11: Examples of Aboriginal and Torres Strait Islander Peoples' practices of adaptation to a changing climate

    `Caring for Country': Traditional Practices for Holistic Land and Cultural Protection and     Source
    Adaptation in a Changing Climate

    Indigenous Protected Area (IPA) management plans enable culturally and ecologically           (Mackey and Claudie,
    compatible development that contribute to local Indigenous economies                          2015).

    IPAs can avoid the potential for `nature­cultures dualism' that locks out Indigenous access   (Lee, 2016)
    in some protected area legislation, as they are based on relational values informed by local
    Indigenous Knowledge

    Fire management using cultural practices can achieve greenhouse gas emission targets while (Robinson et al., 2016)
    also maintaining Indigenous cultural heritage.

    Indigenous Ranger programmes provide a means for Indigenous-guided land management,           (Mackey and Claudie,
    including for fire management and carbon abatement, fauna studies, medicinal plant            2015)
    products, weed management and recovery of threatened species

    Do Not Cite, Quote or Distribute  11-71                                                       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

Faunal field surveys can engage local, bounded and fine-scale intuitive species location by (Wohling, 2009;

Indigenous knowledge holders and their knowledge used for conservation planning  Ziembicki et al., 2013)

          Cultural flows in waterways are a demonstration of cultural knowledge, values and practice (Bark et al., 2015;
          in action as they are informed by Indigenous knowledge, bound by water-dependent values, Taylor et al., 2017)
          and define when and where water is to be delivered - particularly in a changing climate.

 1

 2

 3 11.4.2 Tangata Whenua ­ New Zealand Mori

 4

 5 Mori society faces diverse impacts, risks and opportunities from climate change (Table 11.12). Studies
 6 exploring climate change impacts, scenarios, policy implications, adaptation options and tools for Mori
 7 society have increased substantially e.g. (King et al., 2012; Bargh et al., 2014; Jones et al., 2014; Bryant et
 8 al., 2017; Awatere et al., 2018; Colliar and Blackett, 2018). Mori priorities surrounding climate change
 9 risks and natural resource management have been articulated in planning documents by many Mori kin-
10 groups e.g. (Ngti Tahu- Ngti Whaoa Rnanga Trust, 2013; Raukawa Settlement Trust, 2015; Ngai-Tahu,
11 2018; Te Urunga Kea - Te Arawa Climate Change Working Group, 2021) reflecting the importance of
12 reducing vulnerability and enhancing resilience to climate impacts and risks through adaptation and
13 mitigation.

14

15 Mori have long-term interests in land and water and are heavily invested in climate sensitive sectors
16 (agriculture, forestry, fishing, tourism and renewable energy) (King et al., 2010). Large proportions of
17 collectively owned land already suffer from high rates of erosion (Warmenhoven et al., 2014; Awatere et al.,
18 2018) which are projected to be exacerbated by climate change induced extreme rainfalls (RSNZ, 2016;
19 Awatere et al., 2018) (high confidence). Changing drought occurrence, particularly across eastern and
20 northern New Zealand, is also projected to affect primary sector operations and production (King et al.,
21 2010; Smith et al., 2017; Awatere et al., 2018) (medium confidence). Further, many Mori-owned lands and
22 cultural assets such as marae and urupa are located on coastal lowlands vulnerable to sea-level rise impacts
23 (Manning et al., 2014; Hardy et al., 2019) (high confidence). Mori tribal investment in fisheries and
24 aquaculture faces substantial risks from changes in ocean temperature and acidification, and the downstream
25 impacts for species distribution, productivity and yields (Law et al., 2016) (medium confidence). A clearer
26 understanding of climate change risks and the implications for sustainable outcomes can enable more
27 informed decisions by tribal organisations and governance groups.

28

29 Changing climate conditions are projected to exacerbate health inequities faced by Mori (Bennett et al.,
30 2014; Jones et al., 2014; Hopkins, 2015) (medium confidence). The production and ecology of some
31 keystone cultural flora and fauna may be impacted by projected warming temperatures and reductions in
32 rainfall (RSNZ, 2016; Bond et al., 2019; Egan et al., 2020) (medium confidence). Obstruction of access to
33 keystone species is expected to adversely impact customary practice, cultural identity and well-being (Jones
34 et al., 2014; Bond et al., 2019)(medium confidence). Social-cultural networks and conventions that promote
35 collective action and mutual support are central features of many Mori communities, and these practices are
36 invaluable for initiating responses to, and facilitating recovery from, climate stresses and extreme events
37 (King et al., 2011; Hopkins et al., 2015). Mori tribal organisations have a critical role in defining climate
38 risks and policy responses (Bargh et al., 2014; Parsons et al., 2019) as well as entering into strategic
39 partnerships with business, science, research and government to address these risks (Manning et al., 2014;
40 Beall and Brocklesby, 2017; CCATWG, 2017) (high confidence).

41

42 More integrated assessments of climate change impacts, adaptation and socio-economic risk for different
43 Mori groups and communities, in the context of multiple stresses, inequities and different ways of knowing
44 and being (King et al., 2013; Schneider et al., 2017; Henwood, 2019) would assist those striving to evaluate
45 impacts and risks, and how to integrate these assessments into adaptation plans (high confidence). Better
46 understanding of the social, cultural and fiscal implications of sea-level rise is urgent (PCE, 2015; Rouse et
47 al., 2017; Colliar and Blackett, 2018), including what duties local and central Government might have with
48 respect to actively upholding Mori interests under the Treaty of Waitangi (Iorns Magallanes, 2019) (high
49 confidence). Intergenerational approaches to climate change planning will become increasingly important,
50 elevating political discussions about conceptions of rationality, diversity and the rights of non-human entities
51 (Ritchie, 2013; Carter et al., 2018; Ruru, 2018; Munshi et al., 2020) (high confidence).

Do Not Cite, Quote or Distribute  11-72                                          Total pages: 151
   FINAL DRAFT                                         Chapter 11  IPCC WGII Sixth Assessment Report

1

2

3 Table 11.12: Climate-related impacts and risks for Tangata Whenua New Zealand Mori

   Impact                                                Risks

   Changes in drought occurrence and extreme weather     Risks to Mori tribal investment in forestry, agriculture and
   events                                                horticulture sector operations and production, particularly
                                                         across eastern and northern New Zealand (King et al., 2010;
                                                         Awatere et al., 2018; Hardy et al., 2019)(medium confidence)

   Changes in rainfall, temperature, drought, extreme    Risks to potable water supplies (availability and quality) for
   weather events and ongoing sea-level rise             remote Mori populations (RSNZ, 2016; Henwood,
                                                         2019)(medium confidence)

   Changes in rainfall, temperature, drought, extreme    Risks of exacerbating existing inequities (e.g. health,
   weather events and ongoing sea-level rise             economic, education and social services), social cohesion and
                                                         well-being (Bennett et al., 2014; Jones et al., 2014)(medium
                                                         confidence)

   Changes in rainfall regimes and more intense          Risks to the distribution and survival of cultural keystone flora
   drought combined with degradation of lands and        and fauna, as well as cascading risks for Mori customary
   water                                                 practice, cultural identity and well-being (King et al., 2010;
                                                         RSNZ, 2016; Bond et al., 2019)(high confidence)

   Changes in ocean temperature and acidification        Risks to nearshore and ocean species productivity and
                                                         distribution, as well as cascading risks for Mori tribal
                                                         investment in the fisheries and aquaculture sectors (King et al.,
                                                         2010; Law et al., 2016)(medium confidence)

   Sea-level rise induced erosion, flooding and          Risks to Mori-owned coastal lands and economic investment
   saltwater intrusion                                   as well as risks to community wellbeing from displacement of
                                                         individuals, families and communities (Manning et al., 2014;
                                                         Smith et al., 2017; Hardy et al., 2019)(high confidence)

   Sea-level rise induced erosion, inundation and        Risks to Mori cultural heritage as well as cascading risks for
   saltwater intrusion                                   tribal identity and spiritual well-being (King et al., 2010;
                                                         Manning et al., 2014; RSNZ, 2016)(medium confidence)

   Impacts of climate change, adaptation and mitigation  Risks that governments are unable to uphold Mori interests,
   actions                                               values and practices under the Treaty of Waitangi, creating
                                                         new, modern-day breaches of the Treaty of Waitangi (Iorns
                                                         Magallanes, 2019; MfE, 2020a)(high confidence)

 4
 5

 6 11.5 Cross-Sectoral and Cross-Regional Implications

 7

 8 The impacts and adaptation processes described in sections 11.3 and 11.4 are focused on specific sectors,
 9 systems and Indigenous Peoples. Added complexity, risk and adaptation potential stem from cross-sectoral
10 and cross-regional inter-dependencies.

11

12 11.5.1 Cascading, compounding and aggregate impacts

13

14 11.5.1.1 Observed Impacts

15

16 Climate impacts are cascading, compounding and aggregating across sectors and systems due to complex
17 interactions (high confidence) (Pescaroli and Alexander, 2016; Challinor et al., 2018; Zscheischler et al.,
18 2018; Steffen et al., 2019; AghaKouchak et al., 2020; CoA, 2020e; Lawrence et al., 2020b; Simpson et al.,
19 2021) (Box 11.1; Box 11.3; Box 11.4; Box 11.5; Box 11.6). Cascading impacts propagate via
20 interconnections and systemic factors, including supply chains, shared reliance on connected biophysical
21 systems (e.g. water catchments and ecosystems), infrastructure and essential goods and services, and the
22 exercise of governance, leadership, regulation, resources and standard practices (e.g. in planning and

   Do Not Cite, Quote or Distribute                      11-73                        Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 building codes), including lock-in of past decisions and experience (CSIRO, 2018; Lawrence et al., 2020b).
 2 The capacity of critical systems such as Information, Communication and Technology, water infrastructure,
 3 health care, electricity and transport networks are being stretched, with impacts cascading to other systems
 4 and places, exacerbating existing hazards and generating new risks (Cradock-Henry, 2017) (11.3.6;11.3.10;
 5 Box 11.1). Temporal or spatial overlap of hazards (e.g. drought, extreme heat and fire; drought followed by
 6 extreme rainfall) are compounding impacts (Zscheischler et al., 2018) and affecting multiple sectors.

 7

 8 In Australia, extreme events such as heatwaves, droughts, floods, storms and fires have caused deaths and
 9 injuries (Deloitte, 2017a) (11.3.5.1), and affected many households, communities and businesses via impacts
10 on ecosystems, critical infrastructure, essential services, food production, the national economy, valued
11 places and employment. This has created long-lasting impacts (e.g. mental health, homelessness, health
12 incidents and reduced health services) (Brown et al., 2017; Brookfield and Fitzgerald, 2018; Rychetnik et al.,
13 2019) and reduced adaptive capacity (Friel et al., 2014; O'Brien et al., 2014; Ding et al., 2015; CoA, 2020e)
14 (Box 11.1, Box 11.3, 11.3.1-11.3.10).

15

16 In New Zealand, extreme snow, rainfall and wind events have combined to impact road networks, power and
17 water supply, and have impeded interdependent wastewater and stormwater services and business activities
18 (Deloitte, 2019; Lawrence et al., 2020b; MfE, 2020a) (Box 11.4). Community and infrastructure services are
19 periodically disrupted during extreme weather events, triggering impacts from the interdependencies across
20 enterprises and individuals (Glavovic, 2014; Paulik et al., 2021).

21

22 Slow onset climate change impacts have also had cascading and compounding effects. For example,
23 degradation of the Great Barrier Reef by ocean heating, acidification and non-climatic pressures (Marshall et
24 al., 2019), repeated pluvial, fluvial and coastal flooding of some settlements (Paulik et al., 2019a; Paulik et
25 al., 2020), long droughts and water insecurity in rural communities (Tschakert et al., 2017), and the gradual
26 loss of species and ecological communities, have caused substantial ecological, social and economic losses.
27 Indigenous peoples have especially been impacted by multiple and complex losses (Johnson et al., 2021)
28 (11.4).

29

30 11.5.1.2 Projected Impacts

31

32 Cascading, compounding and aggregate impacts are projected to grow due to a concurrent increase in
33 heatwaves, droughts, fires, storms, floods and sea level (high confidence) (CSIRO, 2020; Lawrence et al.,
34 2020b). Urban wastewater, stormwater and water supply systems are particularly vulnerable in New Zealand
35 (Paulik et al., 2019a; Hughes et al., 2021) to pluvial flooding (Box 11.4) and to sea-level rise (Box 11.6),
36 with flow-on effects to settlements, insurance and finance sectors, and governments (Lawrence et al.,
37 2020b). Furthermore, consecutive heavy rainfall events in late summer and autumn, following drought
38 conditions in low-lying modified wetland areas, have implications for the operation of flood control
39 infrastructure as increased rainfall intensity, land subsidence, and sea-level rise compound and result in the
40 retention of floodwaters (Pingram et al., 2021).

41

42 In Australia, the aggregate loss of wealth due to climate-induced reductions in productivity across
43 agriculture, manufacturing and service sectors is projected to exceed A$19 billion by 2030, A$211 billion by
44 2050 and A$4 trillion by 2100 for RCP8.5 (Steffen et al., 2019) (Table 11.13). Projected impacts also
45 cascade across national boundaries via value chains, markets, movement of humans and other organisms,
46 and geopolitics (e.g. migration from near-neighbours as a pathway for adaptation, mobile climate-sensitive
47 diseases and changes in production and trade patterns) (Lee et al., 2018; Nalau and Handmer, 2018;
48 Schwerdtle et al., 2018; Dellink et al., 2019). The scale of impacts is projected to challenge the adaptive
49 capacity of sectors, governments and institutions (Steffen et al., 2019), including the insurability of assets
50 and risks to lenders (Storey and Noy, 2017).

51

52 11.5.1.3 Adaptation

53

54 Coordinating adaptation strategies and addressing underlying exposure and vulnerability can increase
55 resilience to cascading, compounding and aggregate impacts (Table 11.17; 11.7.3) (high confidence).
56 Systems understanding, network analysis, stress testing, spatial mapping, collaboration, information sharing
57 and interoperability across states, sectors, agencies and value chains, as well as national scale facilitation,

Do Not Cite, Quote or Distribute  11-74       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 can increase adaptive capacity (Espada et al., 2015; CoA, 2020e; Cradock-Henry et al., 2020b; Jozaei et al.,
 2 2020). Greater system diversity, modularity, redundancy, adaptability and decentralised control can reduce
 3 the risk of cascading failures and system breakdown (Sinclair et al., 2017; Sellberg et al., 2018). Addressing
 4 existing vulnerabilities in systems can reduce susceptibility and improve the resilience of interdependent
 5 systems (11.7.3). Multi-level leadership, including national and sub-national policies, laws and finance can
 6 reduce and manage aggregate risks supported by the enablers in Table 11.17.

 7

 8 Anticipatory governance and agile decision making can build resilience to cascading, compounding and
 9 aggregate impacts (Boston, 2016; Deloitte, 2016; Steffen et al., 2019; CoA, 2020e; CSIRO, 2020; Lawrence
10 et al., 2020b; MfE, 2020c) (high confidence). There is uncertainty about whether standard integrated
11 assessment models can estimate cascading and compounding impacts across systems and sectors, but
12 systems methodologies and social network analysis hold promise (Stoerk et al., 2018; Cradock-Henry et al.,
13 2020b). Interventions at the landscape, building and individual scale can reduce the negative health effects of
14 current and future extreme heat, if integrated in well-communicated heat action plans with robust
15 surveillance and monitoring (Jay et al., 2021).

16

17 In Australia, the National Disaster Risk Reduction Framework (CoA, 2018b), National Recovery and
18 Resilience Agency, and Australian Climate Service (CoA, 2021) can provide some support for adaptation
19 across multiple sectors. New Zealand has effective partnerships across critical infrastructure through lifelines
20 groups, but organisational silos and lack of stress testing of plans hamper coordinated decision making
21 during crises and for adaptation (Brown et al., 2017; Lawrence et al., 2020b). The New Zealand national risk
22 assessment, national adaptation plan, forthcoming Climate Change Adaptation Act, and monitoring of
23 adaptation progress by the Climate Change Commission, provide a framework for anticipating climate
24 change risks (MfE, 2020a).

25

26 11.5.2 Implications for National Economies

27

28 The implications of climate change for national economies are significant (high confidence). The costs
29 associated with lost productivity, disaster relief expenditure and unfunded contingent liabilities represent a
30 major risk to financial system stability (MfE, 2020a). Costs include significant and often long-term social
31 impacts, temporary dislocation, business disruption, and impacts on employment, education, community
32 networks, health and wellbeing (Deloitte, 2017a). Climate change disrupts international patterns of
33 agricultural production and trade in ways that may be negative, but may also lead to new opportunities for
34 agriculture (Mosnier et al., 2014; Nelson et al., 2014; Lee et al., 2018). Net exports may increase following
35 global climate shocks (Lee et al., 2018), but the longer term effects on GDP are likely to be negative (Dellink
36 et al., 2019).

37

38 11.5.2.1 Observed Impacts

39

40 In Australia, during 2007-2016, total economic costs from natural disasters averaged A$18.2 billion per year
41 (Deloitte, 2017a). Individual weather-related disaster costs across multiple sectors have exceeded A$4
42 billion, such as the 2009 fires in Victoria (Parliament of Victoria, 2010), the 2010-2011 floods in south-east
43 Queensland (Deloitte, 2017b), the 2019 floods in northern Queensland (Deloitte, 2019) and the 2019-2020
44 fires in southern and eastern Australia (Box 11.1).

45

46 In New Zealand, the annual cost of rural fire to the economy has been estimated at NZ$67 million, with
47 indirect `costs' potentially 2­3 times direct costs (Scion, 2018). Insured losses from weather-related disasters
48 cost almost NZ$1 billion during 2015-2021 (ICNZ, 2021). Floods cost the New Zealand economy at least
49 NZ$120 million for privately insured damages between 2007 and 2017 (D. Frame et al., 2018). The 2007/08
50 drought cost NZ$3.2 billion and the 2012/13 drought cost NZ$1.6 billion, of which about 20% could be
51 attributed to anthropogenic climate change (Frame et al., 2020) (11.5.3.1).

52

53 The intangible costs of climate impacts - including death and injury, impacts on health and wellbeing,
54 education and employment, community connectedness, and the loss of ancestral lands, cultural sites and
55 ecosystems (Barnett et al., 2016; Warner et al., 2019) - affect multiple sectors and systems and exacerbate
56 existing vulnerabilities. While often incommensurable, intangible costs may be far higher than the tangible
57 costs. For example, following the Victorian fires in 2009, the tangible costs were A$3.1 billion while the

Do Not Cite, Quote or Distribute  11-75       Total pages: 151
FINAL DRAFT                       Chapter 11                                        IPCC WGII Sixth Assessment Report

 1 intangible costs were A$3.4 billion; following the Queensland floods in 2010/11, the tangible costs were
 2 A$6.7 billion while the intangible costs were A$7.4 billion (Deloitte, 2016).

 3

 4 11.5.2.2 Projected Impacts

 5

 6 The economic impact increases with higher levels of warming (high confidence) but there is a wide range in
 7 projections. Conservative estimates for the impacts of a 1, 2 or 3°C global warming (relative to 1986-2005)
 8 on Australian GDP growth are -0.3%/year, -0.6%/year and -1.1%/year, respectively, while for New Zealand
 9 the estimates are -0.1%, -0.4%/year and -0.8%/year, respectively (Kompas et al., 2018). More detailed
10 modelling indicates a loss in Australia's GDP of 6% by 2070 for 3°C global warming, while a 2.6% GDP rise
11 by 2070 is possible for 1.5°C global warming (Deloitte, 2020). The potential for much more severe effects on
12 GDP is shown in recent estimates which attempt to account for the increased severity of uncertain effects
13 (e.g. up to 18.5% reduction in Australia's GDP by mid-Century) (Swiss Re, 2021).

14

15 In Australia, the total annual cost of damage due to floods, coastal inundation, forest fires, subsidence and
16 wind (excluding cyclones) is estimated to increase 55% between 2020 and 2100 for RCP8.5 (Mallon et al.,
17 2019). National damage costs and impacts on asset values could be significant (Table 11.13). The macro-
18 economic shocks induced from climate change, including reduced agricultural yields, damage to property
19 and infrastructure and commodity price increases, could lead to significant market corrections and potential
20 financial instability (Steffen et al., 2019). Under a `slow decline' scenario by 2060 where Australia fails to
21 adequately address climate change and sustainability challenges, GDP is projected to grow at 0.7% less per
22 year and real wages would be 50% lower than under an `outlook scenario' where Australia meets climate
23 change and sustainability challenges (CSIRO, 2019).

24

25 In New Zealand, the value of buildings exposed to coastal inundation could increase by NZ$2.55 billion for
26 every 0.1 m increment in sea level, i.e. $25.5 billion for a 1.0 m sea-level rise (Paulik et al., 2020). Greater
27 understanding is required of the distributional impacts, the rate of change of costs over time and the
28 economic implications of delayed action (Warner et al., 2020).

29

30

31 Table 11.13: Economy-wide projected costs (A$) of climate change in Australia. (Estimates are not comparable across
32 studies because different methods have been used. Estimates for later in the century are speculative as both impacts and
33 adaptation are uncertain).

Impact                                                   2030              2050     2090      Reference

Damage-related loss of property value in Australia $571                    $611     $770 billion (Steffen et al., 2019)
                                                                  billion  billion

Property damage in Australia                                               $91 billion $117 billion (Steffen et al., 2019)
                                                                           per year per year
Loss of asset value of road infrastructure (including
freeways, main roads and unsealed roads) in                                         $46-60    (DCCEE, 2011)
Australia at risk of a sea-level rise of 1.1 metres by                              billion
2100
                                                                                    $4.9-6.4  (DCCEE, 2011)
Loss of asset value of rail and tramway                                             billion   (DCCEE, 2011)
infrastructure in Australia at risk of a sea-level rise                                       (DCCEE, 2011)
of 1.1 metres by 2100                                                               $51-72
                                                                                    billion
Loss of asset value of residential buildings in
Australia at risk of a sea-level rise of 1.1 metres by                              $4.2-6.7
2100 (2008 replacement value)                                                       billion

Loss of asset value of light industrial buildings (used
for warehousing, manufacturing, and assembly
activities and services) in Australia at risk of a sea-
level rise of 1.1 metres by 2100

Do Not Cite, Quote or Distribute                         11-76                                Total pages: 151
FINAL DRAFT                                              Chapter 11           IPCC WGII Sixth Assessment Report

Loss of asset value of commercial buildings (used                             $58-81   (DCCEE, 2011)
for wholesale, retail, office and transport activities)                       billion
in Australia at risk of a sea-level rise of 1.1 metres
by 2100 (2008 replacement value)

Accumulated loss of wealth due to reduced                $19         $211     $4.2 trillion (Steffen et al., 2019)
agricultural productivity and labour productivity        billion     billion

Wind damage to dwellings in Cairns, Townsville, $3.8                 $9.7     $20 billion (Stewart and Wang,
Rockhampton and south-east Queensland (assuming billion              billion                    2011)
a 4 per cent discount rate)

Damage to Australian coastal residential buildings                            $8 billion (Wang et al., 2016)
due to sea-level rise (A1B scenario, 3.5°C global
warming)

 1

 2

 3 11.5.2.3 Adaptation

 4

 5 Investments in mitigation and adaptation can help reduce or prevent economic losses now and in the coming
 6 decades (IPCC, 2018; Steffen et al., 2019), however the costs and the benefits of mitigation and adaptation
 7 are not well understood in the region (CSIRO, 2019; MfE, 2020a) (high confidence).

 8

 9 In New Zealand, the emphasis has been on rebuilding after climate disasters, rather than anticipatory
10 adaptation (Boston and Lawrence, 2018). Australia is similarly focused on disaster response and recovery,
11 even though investment in disaster resilience can provide a cost:benefit ratio of 1:2 to 1:11 through reduced
12 post-disaster recovery and reconstruction (GCA, 2019). Recent Australian and state government spending on
13 direct recovery from disasters was around A$2.75 billion per year, compared to funding for natural disaster
14 resilience of approximately A$0.1 billion per year (Deloitte, 2017b). The Australian Government is
15 supporting most of the 80 recommendations from the Royal Commission into National Natural Disaster
16 Arrangements, including establishing a disaster advisory body and a resilience and recovery agency (CoA,
17 2020e; CoA, 2020b). Australia and New Zealand provide humanitarian and disaster assistance across the
18 Pacific, which is increasingly focused on climate adaptation and the Sustainable Development Goals (Brolan
19 et al., 2019) as cyclones and floods become amplified by climate change (Fletcher et al., 2013) (Table 11.3).
20 Climate change may increase current migration flows to and impacts on diaspora in Australia and New
21 Zealand from near neighbour island nations, as they become increasingly stressed by rising seas, higher
22 temperatures, more droughts and stronger storms (Nalau and Handmer, 2018).

23

24 Delaying adaptation to climate risks may result in higher overall costs in future when adaptation is more
25 urgent and impacts more extreme (Boston and Lawrence, 2018; IPCC, 2018) (medium confidence).
26 Estimates of the magnitude of adaptation costs and benefits in the region are localised and sectoral, e.g.
27 (Thamo et al., 2017) or regionally aggregated (Joshi et al., 2016). Adaptation costs are expected to increase
28 markedly for higher RCPs, e.g. a tripling of expected costs between RCP2.6 and RCP8.5 for sea-level rise
29 protection in Australia (Ware et al., 2020). Existing governance arrangements for funding adaptation are
30 inadequate for the scope and scale of climate change impacts anticipated; dedicated funding mechanisms that
31 can be sustained over generations can enable more timely adaptation (Boston and Lawrence, 2018).

32

33

34 11.6 Key Risks and Benefits

35

36 Nine key risks have been identified (Table 11.14) based on four criteria: magnitude, likelihood, timing and
37 adaptive capacity (Chapter 16). Most of the key risks are similar to those in the IPCC AR5 Australasia
38 chapter (Reisinger et al., 2014), but the emphasis here is on specific systems affected by multiple hazards
39 rather than specific hazards affecting multiple systems. The selection of key risks reflects what has been
40 observed, projected and documented, noting that there are gaps in knowledge, and a lack of knowledge does
41 not imply a lack of risk (11.7.3.3). Key risks are grouped into four categories:

42

Do Not Cite, Quote or Distribute                         11-77                         Total pages: 151
    FINAL DRAFT                               Chapter 11  IPCC WGII Sixth Assessment Report

1 Ecosystems at critical thresholds where recent climate change has caused significant damage and further

2 climate change may cause irreversible damage, with limited scope for adaptation

3   1. Loss and degradation of coral reefs in Australia and associated biodiversity and ecosystem service

4   values due to ocean warming and marine heatwaves (11.3.2.1, 11.3.2.2, Box 11.2).

5   2. Loss of alpine biodiversity in Australia due to less snow (11.3.1.1, 11.3.1.2).

6

7 Key risks that have potential to be severe but can be reduced substantially by rapid, large-scale and effective

8 mitigation and adaptation

9   3. Transition or collapse of alpine ash, snowgum woodland, pencil pine and northern jarrah forests in

10  southern Australia due to hotter and drier conditions with more fires (11.3.1.1, 11.3.1.2)

11  4. Loss of kelp forests in southern Australia and southeast New Zealand due to ocean warming, marine

12  heatwaves and overgrazing by climate-driven range extensions of herbivore fish and urchins

13  (11.3.2.1, 11.3.2.2).

14  5. Loss of natural and human systems in low-lying coastal areas due to sea level rise (11.3.5, Box

15  11.6).

16  6. Disruption and decline in agricultural production and increased stress in rural communities in south-

17  western, southern and eastern mainland Australia due to hotter and drier conditions (11.3.4, 11.3.5,

18  Box 11.3).

19  7. Increase in heat-related mortality and morbidity for people and wildlife in Australia due to

20  heatwaves (11.3.5.1, 11.3.5.2, 11.3.6.1, 11.3.6.2).

21

22 Key cross-sectoral and system-wide risk

23  8. Cascading, compounding and aggregate impacts on cities, settlements, infrastructure, supply-chains

24  and services due to wildfires, floods, droughts, heatwaves, storms and sea-level rise (11.5.1.1,

25  11.5.1.2, Box 11.1, Box 11.4, Box 11.6).

26

27 Key implementation risk

28  9. Inability of institutions and governance systems to manage climate risks. (11.5; 11.7.1, 11.7.2,

29  11.7.3).

30

31 At higher levels of global warming, adaptation costs increase, options become limited and risks grow. The
32 `burning embers' diagram in Figure 11.6 has four IPCC risk categories: "undetectable", "moderate", "high"
33 and "very high", with transition points defined by different global warming ranges. The embers are

34 indicative, based on an assessment of available literature and expert judgement (Supplementary Material SM
35 11.2). Outcomes for low and moderate adaptation have been compared, with the latter including both
36 incremental and transformative options. Illustrative examples of adaptation pathways are shown in Figure
37 11.7 for low-lying coastal areas and Figure 11.8 for heat-related mortality. These figures highlight thresholds

38 at which adaptation options become ineffective, and possible combinations of strategies and options

39 implemented at different times to manage emerging risks and changing risk profiles.

40

41 Caveats: (a) key risks are assessed at regional scales, so they do not include other risks for finer scales or

42 specific groups; (b) non-climatic vulnerabilities are held constant for simplicity; (c) the assessment of risk

43 ratings at different levels of global warming is limited by available literature; (d) risks increase with global
44 warming, despite the lack of an IPCC risk rating beyond "very high"; and (e) the feasibility and effectiveness
45 of adaptations options were not assessed due to limited literature (11.7.3.3).

46

47 The New Zealand National Climate Change Risk Assessment (MfE, 2020a) identified the priority risks from
48 climate change for New Zealand based on a literature review and expert elicitation. The top two risks in each
49 of five domains are: Natural environment (1) risks to coastal ecosystems due to ongoing sea-level rise and
50 extreme weather events, (2) risks to indigenous ecosystems and species from invasive species; Human
51 environment (1) risks to social cohesion and community well-being from displacement of people, (2) risks of
52 exacerbating existing inequities and creating new and additional inequities from distribution impacts;
53 Economy (1) risks to governments from economic costs associated with lost productivity, disaster relief
54 expenditure and unfunded contingent liabilities, (2) risks to the financial system from instability; Built

55 environment (1) risk to potable water supplies due to changes in rainfall, temperature, drought, extreme

56 weather events and ongoing sea-level rise,(2) risks to buildings due to extreme weather events, drought,
57 increased fire weather and ongoing sea-level rise; Governance (1) risk of maladaptation due to practices,

    Do Not Cite, Quote or Distribute          11-78                                     Total pages: 151
FINAL DRAFT                       Chapter 11                   IPCC WGII Sixth Assessment Report

 1 processes and tools that do not account for uncertainty and change over long timeframes, and (2) risk that
 2 climate change impacts across all domains will be exacerbated, because current institutional arrangements
 3 are not fit for adaptation. Not all of these risks feature as key risks for the wider Australasia region;
 4 nonetheless they are reflected across Chapter 11 and remain priorities for New Zealand to address through
 5 the National Adaptation Plan, its implementation and monitoring.

 6

 7 Short-term benefits from climate change may include reduced winter mortality, reduced energy demand for
 8 winter heating, increased agriculture productivity and forest growth in south and west New Zealand, and
 9 increased forest and pasture growth in southern Australia except where rainfall and soil nutrients are limiting
10 (11.3.4; 11.3.6; 11.3.10) (medium confidence).

11

12

13 Table 11.14: Key risks from climate change based on assessment of the literature and expert judgement
14 (Supplementary Material SM 11.2). Assessment criteria are magnitude, timing, likelihood and adaptive capacity. Risk
15 drivers are hazards, exposure and vulnerability. Adaptation options describe ways in which risks can be reduced.
16 Confidence ratings are based on the amount of evidence and agreement between lines of evidence.

Key risk               Consequences influenced by hazards, exposure, vulnerability and adaptation options

(confidence rating)
(Chapter reference)

1. Loss and            Consequences: Widespread destruction of coral reef ecosystems and dependent socio-

degradation of         ecological systems. Three mass bleaching events from 2016-2020 have already caused

tropical shallow coral significant loss of corals in shallow-water habitats across the Great Barrier Reef. Globally,

reefs and associated bleaching is projected to occur twice each decade from 2035 and annually after 2044 under

biodiversity and       RCP 8.5 and annually after 2051 under RCP4.5. A 3oC global warming could cause over

ecosystem service six times the 2016 level of thermal stress.

values in Australia    Hazards: Increase in background warming and marine heatwave events degrade reef-
due to ocean warming   building corals by triggering coral bleaching events at a frequency greater than the recovery
and marine             time. Fish populations also decline during and following heat wave events.
heatwaves

                       Exposure: Increasing geographic area affected by rate and severity of ocean warming

(very high             Vulnerability: Vulnerability to increases in sea temperature is already very high because of
confidence)            other stressors on the ecosystem, including sediment, pollutants, and overfishing.

(11.3.2, Box 11.2)     Adaptation options: Minimising other stressors. Efforts on the Great Barrier Reef may slow
                       the impacts of climate change in small sections or reduce short-term socio-economic
                       ramifications, but will not prevent widespread bleaching.

2. Loss of alpine      Consequences: Loss of endemic and obligate alpine wildlife species and plant communities
biodiversity in        (feldmark and short alpine herb-fields) as well as increased stress on snow-dependent plant
Australia due to less  and animal species.
snow
                       Hazards: Projected decline in annual maximum snow depth by 2050 is 30-70% (low
(high confidence)      emissions) and 45-90% (high emissions); projected increases in temperature and decreases
                       in precipitation.
(11.3.1, Tables 11.2,
11.3, 11.4, 11.5)      Exposure: Alpine species face elevation squeeze due to lack of nival zone and alpine
                       environments have restricted geographic extent.

                       Vulnerability: Narrow ecological niche of species including snow-related habitat
                       requirements; encroachment from sub-Alpine woody shrubs; vulnerability generated by
                       non-climatic stressors including weeds and feral animals, especially horses

                       Adaptation options: Reducing pressure on alpine biodiversity from land uses that degrade
                       vegetation and ecological condition, along with weed and pest management.

Do Not Cite, Quote or Distribute  11-79                        Total pages: 151
FINAL DRAFT                                       Chapter 11  IPCC WGII Sixth Assessment Report

3. Transition or         Consequences: If regenerative capacities of the dominant (framework) canopy tree species

collapse of alpine are exceeded, a long lasting or irreversible transition to a new ecosystem state is projected

ash, snowgum             with loss of characteristic and framework species including loss of some narrow range

woodland, pencil pine endemics.

and northern jarrah      Hazards: Hotter and drier conditions have increased extreme fire weather risk since 1950,
forests in southern      especially in southern and eastern Australia. The number of severe fire weather days is
Australia due to         projected to increase 5-35% (RCP2.6) and 10-70% (RCP8.5) by 2050
hotter and drier

conditions with more Exposure: Shift in landscape fire regimes to larger, more intense and frequent wildfires

fires                    over extensive areas (~10 million hectares) of forests and woodlands from longer fire

                         seasons and more hazardous fire conditions and increasing human-sourced ignitions from

                         urbanisation and projected increase in frequency of lightning strikes

(high confidence)        Vulnerability: The resilience and adaptive capacity of the forests is being reduced by

                         ongoing land clearing and degrading land management practices

(11.2, 11.3.1, 11.3.2, Adaptation options: Increased capacity to extinguish wildfires during extreme fire weather

Box 11.1)                conditions; avoiding and reducing forest degradation from inappropriate forest management

                         practices and land use.

4. Loss of kelp forests  Consequences: Observed decline in giant kelp in Tasmania since 1990, with less than 10%
in southern Australia    remaining by 2011 due to ocean warming. Extensive loss of kelp -140,187 hectares across
and southeast New        Australia. Loss of bull kelp in southern New Zealand, replaced by the introduced kelp
Zealand due to ocean     following the 2017/18 marine heatwave. Further loss of native kelp is projected with
warming, marine          warming oceans.
heatwaves and
overgrazing by           Hazards: Ocean warming and marine heatwave events
climate-driven range
extensions of            Exposure: Coastal waters around Australia and New Zealand
herbivore fish and
urchins                  Vulnerability: Giant kelp are already Federally listed in Australia as an endangered marine
                         community type. In Australia, kelp forests are vulnerable to nutrient poor East Australian
                         Current waters pushing further south, warming waters and increased herbivory from range-
                         extending species.

(high confidence)        Adaptation options: Minimizing other stressors, local restoration, and transplantation of
(11.3.2)                 heat-tolerant phenotypes.

5. Loss of human and     Consequences: Nuisance and extreme coastal flooding are already occurring due to sea-
natural systems in       level rise (SLR). For 0.2-0.3 m SLR, coastal flooding is projected to become more frequent,
low-lying coastal        e.g. current 1-in-100 year flood would occur every year in Wellington and Christchurch.
areas from ongoing       For 0.5 m SLR, the value of buildings in New Zealand exposed to coastal inundation could
sea-level rise           increase by NZ$12.75 billion and the current 1-in-100 year flood in Australia could occur
                         several times a year. For 1.0 m SLR, the value of exposed assets in New Zealand would be
(high confidence)        NZ$25.5 billion. For 1.1 m SLR, the value of exposed assets in Australia would be A$164-
                         226 billion. This would be associated with displacement of people, disruption and reduced
(11.2, 11.3.2, 11.3.5,   social cohesion, degraded ecosystems, loss of cultural heritage and livelihoods, and loss of
11.3.10, 11.4, Table     traditional lands and sacred sites.
11.3; Box 11.6)
                         Hazards: Rising sea level (0.2-0.3 m by 2050, 0.4-0.7 m by 2090), storm surges, rising
                         ground water tables.

                         Exposure: Population growth, new and infill urbanization, tourism developments in low-
                         lying coastal areas. Buildings, roads, railways, electricity and water infrastructure. Torres
                         Strait Island and remote Mori communities are particularly exposed and sensitive.

                         Vulnerability: Ineffective planning regulations, reduced availability and increased cost of
                         insurance, and costs to governments as insurers of last resort. Inadequate investment in
                         avoidance and preparedness exacerbating underlying social vulnerabilities. Financial and
                         physical capacities to cope and adapt are uneven across populations, creating equity issues.

                         Adaptation options: Risk reduction coordinated across all levels of government with
                         communities. Statutory planning frameworks, decision tools and funding mechanisms that
                         can address the changing risk. Planning and land use decisions, including managed retreat
                         where it is inevitable. Improved capacity of emergency services, early warning systems,

Do Not Cite, Quote or Distribute                  11-80                                         Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

                         improved planning and regulatory practice and building and infrastructure design standards.
                         Options that anticipate risk and adjust as conditions change.

6. Disruption and        Consequences: Projected decline in crop, horticulture and dairy production. e.g. decline in
decline in agricultural  median wheat yields by 2050 of up to 30% in south-west Australia and up to 15% in South
production and           Australia. Increased heat stress in livestock by 31­42 days per year by 2050. Reduced
increased stress in      winter chilling for horticulture. Increased smoke impacts for viticulture. Flow-on effects for
rural communities        agricultural supply chains, farming families and rural communities across south-western,
across south western,    southern and south-eastern Australia, including the Murray-Darling Basin (MDB).
southern and eastern
mainland Australia       Hazards: Hotter and drier conditions with constraints on water resources and more frequent
due to hotter and        and severe droughts in south-western, southern and eastern Australia.
drier conditions.
                         Exposure: Across south western, southern and eastern Australia, many production regions
(high confidence)        are exposed including the MDB which supports agriculture worth A$24 billion/year, 2.6
                         million people in diverse rural communities, and important environmental assets containing
                         16 Ramsar listed wetlands.

(11.2, 11.3.4,           Vulnerability: Existing financial, social, health and environmental pressures on rural,
11.3.6.3, 11.4.1,        regional and remote communities. Existing competition for water resources among
Table 11.11, Boxes       communities, industries and environment, and uncertainty about sharing of water under a
11.1, 11.3)              drying climate.

                         Adaptation options: Improved governance and collaboration to build rural resilience,
                         including regional and basin-scale initiatives. Improved water policies and initiatives (e.g.
                         MDB Plan) and changes in management and technologies. Resilience-focused planning for
                         rural settlements, land-use, industry, infrastructure and value chains. Adoption of
                         information, tools and methods to better manage uncertainty, variability and change.
                         Incremental changes in farm management practices (e.g. stubble retention, weed control,
                         water-use efficiency, sowing dates, cultivars). In some regions, major changes may be
                         necessary, e.g. diversification in agricultural enterprises, transition to different land-uses
                         (e.g. carbon sequestration, renewable energy production, biodiversity conservation) or
                         migration to another area. Flows in waterways based on Indigenous knowledge to protect
                         cultural assets.

7. Increase in heat-     Consequences: During 1987-2016, natural disasters caused 971 deaths and 4,370 injuries,
related mortality and    with more than 50% due to heatwaves. Annual increases are projected for excess deaths,
morbidity for people     additional hospitalisations and ambulance callouts. Heatwave related excess deaths in
and wildlife in          Melbourne, Sydney and Brisbane are projected to increase by about 300/year (RCP2.6) to
Australia                600/year (RCP8.5) during 2031-2080 relative to 142/year during 1971-2020, assuming no
                         adaptation. Significant heat-related mortality of wildlife species (flying foxes, freshwater
(high confidence)        fish) has been observed and is projected to increase.

(11.2, 11.3.1, 11.3.5,   Hazards: Increased frequency, intensity and duration of extreme heat events
11.3.6, 11.4)
                         Exposure: Pervasive, but differentially affecting some wildlife species depending on their
                         thermal tolerances and occupational groups (e.g. outdoor workers) and those living in high
                         exposure areas (e.g. urban heat islands). Health risks multiply with other harmful
                         exposures, e.g. to wildfire smoke.

                         Vulnerability: Lower adaptive capacity for young/old/sick people, those in low quality
                         housing and lower socio-economic status, and areas served by fragile utilities (power,
                         water). Remote locations with extreme heat and inadequate cooling in housing
                         infrastructure (such as remote indigenous communities). For wildlife, impacts of extreme
                         heat events are being amplified by habitat loss and degradation.

                         Adaptation options: Urban cooling interventions including irrigated green infrastructure
                         and increased albedo, education to reduce heat stress, heatwave/fire early-warning systems,
                         battery/generator systems for energy system security, building standards that improve
                         insulation/cooling, accessible / well-resourced primary health care. For wildlife, removing
                         human stressors, reducing pressures from ferals and weeds, and ensuring there is suitable
                         habitat.

8. Cascading,            Consequences: Widespread and pervasive damage and disruption to human activities
compounding and          generated by interdependencies and interconnectedness of physical, social and natural

Do Not Cite, Quote or Distribute  11-81       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

aggregate impacts on     systems. Examples include: Failure of transport, energy and communication infrastructure
cities, settlements,     and services, heat-stress, injuries and deaths, air pollution, stress on hospital services,
infrastructure,          damage to agriculture and tourism, insurance loss from heatwaves and fires; failure of
supply-chains and        transport, stormwater and flood-control infrastructure and services from floods and storms;
services due to          water restrictions, reduced agricultural production, stress for rural communities, mental
extreme events           health issues, lack of potable water from droughts; damage to buildings, roads, railways,
                         electricity and water infrastructure, loss of assets and lives, displacement of people, reduced
(high confidence)        social cohesion, and degraded ecosystems from extreme sea-level rise. Large aggregate
                         costs due to lost productivity and major disaster relief expenditure, creating unfunded
(11.2, 11.3.4, 11.3.5,   liabilities and supply chain disruption, e.g., the 2019-2020 Australian fires cost A$8 billion.
11.3.6, 11.3.7, 11.3.8,  The impact of a 1, 2 or 3°C global warming (relative to 1986-2005) on Australian GDP
11.3.9, 11.3.10, 11.4,   growth is estimated at -0.3%/year, -0.6%/year and -1.1%/year, respectively, while for New
11.5.1, Boxes 11.1,      Zealand estimates are -0.1%/year, -0.4%/year and -0.8%/year, respectively. Impacts on
11.4, 11.6)              Mori tribal investments in forestry, agriculture, horticulture, fisheries and aquaculture.

                         Hazards: Heatwaves, droughts, fires, floods, storms and sea-level rise. This includes
                         cascading and compound events such as heatwaves with fires, storms with floods, or
                         droughts followed by heavy rainfall and extreme sea levels.

                         Exposure: Highly populated areas, rural and remote settlements, traditional lands and
                         sacred sites. Greater urban density and population growth increases exposure in high-risk
                         areas. Different exposure for different hazards, e.g. heatwaves: urban and peri-urban areas;
                         fire: peri-urban areas and settlements near forests; floods: people, property and
                         infrastructure from pluvial floods in cities and settlements and fluvial floods on floodplains;
                         storms: buildings and infrastructure in cities and settlements.

                         Vulnerability: Existing social and economic challenges (e.g. those caused by COVID-19)
                         and socio-economic and cultural inequalities; competing resource and land use demands
                         across sectors; inadequate planning, policy, governance, decision making and disaster
                         resilience capacity; and non-climatic stresses on ecosystems. Vulnerabilities generated by
                         interdependencies and interconnectedness of physical, social and natural systems.

                         Adaptation options: Flexible and timely adaptation strategies that prepare socio-economic
                         and natural systems for surprises and unexpected threats. Multi-sector coordinated actions
                         that address widespread impacts, redress existing vulnerabilities and building adaptive
                         capacity and systemic resilience. Improved coordination between and within levels of
                         governments, communities and private sector. Greater use of dynamic decision frameworks
                         and suitable economic and social assessment tools. Improved emergency services and early
                         warning systems; use of climate resilient standards for buildings and infrastructure.
                         Transformational adaptations e.g. managed retreat, that can be planned in stages.

9. Inability of          Consequences: Climate hazards overwhelm the capacity of institutions, organisations,
institutions and         systems and leaders to provide necessary policies, services, resources, coordination and
governance systems       leadership. Failed adaptation at the institutional and governance level has widespread,
to manage climate        pervasive impacts for all areas of society. This includes a reliance on reactive, short-term
risks                    decision making that locks in existing exposures, leaves perverse incentives and
                         interconnected and systemic impacts unaddressed, and generates high costs, fiscal impacts.
(high confidence)        This worsens vulnerability and leads to maladaptation, inequities and injustices within and
                         across generations, as well as actions that do not uphold the rights, interests, values and
(11.2, 11.3.5, 11.3.6,   practices of Indigenous Peoples. Resultant failure to take adaptation action generates
11.3.7, 11.3.8,          litigation risk.
11.3.10, 11.4, 11.5.1,
11.7.2, Boxes 11.1-      Hazards: The increasing frequency, duration, severity and complexity of extreme weather
11.6)                    events, droughts and sea-level rise

                         Exposure: All sectors, communities, organisations, and governments

                         Vulnerability: Fragmented institutional and legal arrangements, under-resourcing of
                         services, lack of dedicated adaptation funding instruments and resources to support
                         communities and local government, uneven capability to manage uncertainty, and
                         conflicting values and competing policy and political interests.

                         Adaptation options: Pre-emptive options that avoid and reduce risks. Redesign of policy
                         and statutory frameworks, and funding instruments for addressing changing risks and
                         uncertainties that enable just and collaborative governance across scales and domains.
                         Addressing existing vulnerabilities, and capacity, capability and leadership deficits within

Do Not Cite, Quote or Distribute  11-82       Total pages: 151
      FINAL DRAFT                 Chapter 11  IPCC WGII Sixth Assessment Report

1                  and across all levels of government, all sectors, Indigenous peoples and communities. Risk
2                  and vulnerability assessment methodologies and decision-making tools that build resilience
                   and address changing risks and vulnerabilities. Co-designed adaptation approaches
                   implemented with communities, including Mori tribal organisations and Australian
                   Aboriginal and Torres Strait Island peoples.

Do Not Cite, Quote or Distribute  11-83       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1

2 Figure 11.6: Burning embers diagram for each of the nine key risks for low and moderate adaptation. The risk
3 categories are undetectable, moderate, high and very high. While there is no risk category beyond very high, risks
4 obviously get worse with further global warming, and the risk for coral reefs is already very high. The assessment is
5 based on available literature and expert judgement, summarised in Table 11.14 and described in Supplementary
6 Material SM 11.2. The global warming range associated with each risk transition has a confidence rating (**** very
7 high, *** high, ** moderate, * low) based on the amount of evidence and level of agreement between lines of evidence.

8

Do Not Cite, Quote or Distribute  11-84       Total pages: 151
      FINAL DRAFT                            Chapter 11  IPCC WGII Sixth Assessment Report

1

2

3 Figure 11.7: Illustrative adaptation pathway for risk to natural and human systems in low-lying coastal areas due to
4 sea-level rise.

5

6

7

8 Figure 11.8: Illustrative adaptation pathway for risk of heat-related mortality and morbidity for people and wildlife in

9 Australia due to heatwaves.

10

11

12 11.7 Enabling Adaptation Decision-making

13

14 11.7.1  Observed Adaptation Decision-Making

15

16 The ambition, scope and progress on adaptation by governments has risen, but is uneven with a focus on

17 high level strategies at national level adaptation planning at sub-national levels and new enabling legislation

18 (Tables 11.15a and 11.15b; (Lawrence et al., 2015; Macintosh et al., 2015; MfE, 2020a) (very high

19 confidence). The adaptation process comprises vulnerability and risk assessments, identification of options,
20 planning, implementation, monitoring, evaluation and review. Large gaps remain, especially in effective

    Do Not Cite, Quote or Distribute         11-85       Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 implementation, monitoring and evaluation (Supplementary Material SM 11.1) (CCATWG, 2017; Warnken
 2 and Mosadeghi, 2018) and current adaptation is largely incremental and reactive (Box 11.4, Box 11.6, Table
 3 11.14) (very high confidence).

 4

 5 Australia has a National Climate Resilience and Adaptation Strategy, and a National Recovery and
 6 Resilience Agency (11.5.2.3), the first National Action Plan to implement the Disaster Risk Reduction
 7 Framework acknowledges climate change as a disaster risk driver (Home Affairs, 2020). States and
 8 territories have climate change adaptation strategies with plans to address them (Table 11.15a).with some
 9 adaptation implementation at state level and increasingly at local government level (Jacobs et al., 2016;
10 Warnken and Mosadeghi, 2018) (Table 11.15a). In coastal zones, however, few local government planning
11 instruments are being applied (Warnken and Mosadeghi, 2018; Harvey, 2019; Robb et al., 2019; Elrick-Barr
12 and Smith, 2021). Some businesses and industry sectors are recognizing climate-related risks and adaptation
13 planning (11.3.4; 11.3.7; 11.3.10) (Harris et al., 2016; Hennessy et al., 2016; CBA, 2019). There is an
14 opportunity for Australia to undertake a national risk assessment and to develop a national climate adaptation
15 implementation plan that is aligned with Paris Agreement expectations of a national level system for
16 adaptation planning, monitoring and reporting (Morgan et al., 2019).

17

18 New Zealand's Climate Change Response Act in 2019 creates a legal mandate for National Climate Change
19 Risk Assessments (first one completed) (MfE, 2020a) and National Adaptation Plans (first in preparation)
20 and a Climate Change Commission to monitor and report on adaptation implementation. Preparation of
21 Natural and Built Environment, Strategic Planning and Climate Change Adaptation Acts is underway,
22 including provision for funding and managed retreat (MfE, 2020c). National coastal guidance is available for
23 adaptation planning to address changing climate risks (MfE, 2017a) (Table 11.15b). Meanwhile, several
24 local authorities have developed integrated climate change strategies and plans and revised policies and rules
25 to enable adaptation (Table 11.15b). Different adaptation approaches continue to create confusion and inertia
26 while development pressures continue (Schneider et al., 2017). Opportunities for integrated adaptation and
27 mitigation planning in regional policies and plans have arisen through the Resource Management
28 Amendment Act 2020 (Dickie, 2020), the National Policy Statement on Freshwater Management (MfE,
29 2020b), and the revised national coastal guidance (MfE, 2017a), but rely on funding instruments to be in
30 place and statutes are aligned for their effectiveness (Boston and Lawrence, 2018; CCATWG, 2018) (very
31 high confidence).

32

33 There is a growing awareness of the need for more proactive adaptation planning at multiple scales and
34 across sectors, and a better understanding of future risks and limits to adaptation is emerging (Evans et al.,
35 2014; Archie et al., 2018; Christie et al., 2020; MfE, 2020a) (medium confidence). Disaster risk reduction is
36 being positioned as part of climate change adaptation (Forino et al., 2017; CDEM, 2019; Forino et al., 2019;
37 CoA, 2020e; CSIRO, 2020). Public and private climate adaptation services are informing climate risk
38 assessments, but are characterized by fragmentation, duplication, inconsistencies, poor governance and
39 inadequate funding - addressing these gaps presents adaptation opportunities (CCATWG, 2018; Webb et al.,
40 2019; NESP ESCC, 2020) (Tables 11.15a; 11.15b). Large infrastructure asset planning is starting to factor
41 in climate risks, but implementation is variable (Gibbs, 2020). Local governments in Australia are
42 increasingly implementing adaptation plans but few monitor or evaluate actual outcomes or know how to
43 (Scott and Moloney, 2021).

44

45 Observed and projected rates of sea-level rise (Box 11.6) and increased flood frequency (11.3.3) are
46 challenging established uses of modelling, risk assessment, and cost benefit analysis, where climate change
47 damage functions cannot be projected or are unknown (deep uncertainty), or impacts on communities are
48 ambiguous (Infometrics and PSConsulting, 2015; Lawrence et al., 2019a; MfE, 2020a). New tools are
49 available in the region (Table 11.17) but uptake cannot be assumed (Lawrence and Haasnoot, 2017;
50 Palutikof et al., 2019c) (high confidence).

51

52 Resilience and adaptation approaches are beginning to converge (White and O'Hare, 2014; Aldunce et al.,
53 2015) (Supplementary Material SM 11.1) but widespread "bounce back" resilience-driven responses that
54 lock in risk by discounting ongoing and changing climate risk (Leitch and Bohensky, 2014; O'Hare et al.,
55 2016; Wenger, 2017; Torabi et al., 2018) can create maladaptation and impede long-term adaptation goals
56 (Glavovic and Smith, 2014; Dudney et al., 2018) (high confidence).

57

Do Not Cite, Quote or Distribute  11-86       Total pages: 151
FINAL DRAFT                                 Chapter 11  IPCC WGII Sixth Assessment Report

1 Local government engagement with communities on adaptation is starting to motivate a change towards
2 more collaborative engagement practices (Archie et al., 2018; Bendall, 2018; MfE, 2019; Schneider et al.,
3 2020). Nature-based adaptations (Colloff et al., 2016; Lavorel et al., 2019; Della Bosca and Gillespie, 2020)
4 and `green infrastructure' (Lin et al., 2016; Alexandra and Norman, 2020) are increasingly being adopted
5 (Rogers et al., 2020a) (medium confidence).

 6

 7 Some businesses have initiated active adaptation (Aldum et al., 2014; Linnenluecke et al., 2015; Bremer and
 8 Linnenluecke, 2017; CCATWG, 2017; MfE, 2018) with most focused on identifying climate risks (Aldum et
 9 al., 2014; Gasbarro et al., 2016; Cradock-Henry, 2017). Businesses are more likely to engage in anticipatory
10 adaptation when the frequency of climate events is known (McKnight and Linnenluecke, 2019). Effective
11 cooperation and a positive innovation culture can contribute to the collaborative development of climate
12 change adaptation pathways (Bardsley et al., 2018) (medium confidence).

13

14 Some areas in northern Australia and New Zealand, especially those with higher proportions of Indigenous
15 populations, face severe housing, health, education, employment and services deficits that exacerbate the
16 impacts of climate change (Kotey, 2015) (11.3.5; 11.4; 11.6). Where adaptation relies upon an aging
17 population and an over-stretched volunteer base, vulnerability to climate change impacts is being
18 exacerbated (Astill and Miller, 2018; Davies et al., 2018). Adaptation options that succeed within remote
19 Indigenous communities are founded on connections to traditional lands, alignment with cultural values and
20 contribute to social, cultural and economic goals (Nursey-Bray and Palmer, 2018). Knowledge co-production
21 for Indigenous adaptation pathways can enable transformative change from colonial legacies (Hill et al.,
22 2020). Learning and experimentation across governance boundaries and between agencies and local
23 communities enables adaptation to be better aligned with changing climate risks and community (Fünfgeld,
24 2015; Howes et al., 2015; Bardsley and Wiseman, 2016; Lawrence et al., 2019b) (high confidence).

25

26 There is increasing focus on improving adaptive capacity for transitional and transformational responses, but
27 reactive responses dominate (Smith et al., 2015; Schlosberg et al., 2017; Boston and Lawrence, 2018) (very
28 high confidence). While extreme events can provide opportunities for positive transitions within
29 communities (Cradock-Henry et al., 2018b) (for example the Queensland Reconstruction Authority Building
30 Back Better scheme), often rebuilding occurs in at-risk places to aid quick recovery (Lawrence and
31 Saunders, 2017). Community-based adaptation innovations (Kench et al., 2018; Forino et al., 2019){Bendall,
32 2018 #413} include: relationship building; use of new decision tools, pathways planning with communities,
33 visualisation and serious games (Lawrence and Haasnoot, 2017; Schlosberg et al., 2017; Flood et al., 2018;
34 Reiter et al., 2018; Serrao-Neumann and Choy, 2018); communities of practice; and climate information
35 sharing (Astill et al., 2019; Stone et al., 2019).

36

37

38 Table 11.15a: Examples of Australian adaptation strategies, plans and initiatives by government agencies at the (a)
39 national level, (b) sub-national, and (c) regional or local level. These examples have not been assessed for their
40 effectiveness (see Supplementary Material Table SM11.1a).

Jurisdiction    Strategies /Plans /Actions

National Level  National Climate Resilience and Adaptation Strategy 2015 (CoA, 2015)
Australia       National Disaster Risk Reduction Framework (2018) (CoA, 2018b)
                National Recovery and Resilience Agency and Australian Climate Service (CoA, 2021)

Sub-national

Australian Capital ACT Climate Change Strategy 2019-2025 (ACT Government, 2019)
Territory (ACT) Canberra's Living Infrastructure Plan: Cooling the City (ACT Government, 2020b); ACT

                         Wellbeing Framework (ACT Government, 2020a)

Do Not Cite, Quote or Distribute            11-87       Total pages: 151
FINAL DRAFT                       Chapter 11                            IPCC WGII Sixth Assessment Report

New South Wales NSW Climate Change Policy Framework (NSW Government, 2016)

                 Coastal Management Framework (OEH, 2018b) including:
                 Coastal Management Act 2016; State Environmental Planning Policy (Coastal
                 Management) 2018; NSW Coastal Management Manual (OEH, 2018c; OEH, 2018a)

Northern Territory Northern Territory Climate Change Response: Towards 2050 (DENR, 2020b); Three-year
                         action plan (DENR, 2020a)

Queensland       Pathways to climate resilient Queensland: Queensland Climate Adaptation Strategy 2017-
                 2030 (DEHP, 2013)

                 Sector adaptation plans https://www.qld.gov.au/environment/climate/climate-
                 change/adapting/sectors-systems

                 State heatwave risk assessment 2019 (QFES, 2019)

                 Planning Act 2016 (Queensland Government, 2020) and the Coastal Protection and
                 Management Act 1995 (Queensland Government, 1995) plus supporting initiatives:
                 Coastal Management Plan (DEHP, 2013); Shoreline Erosion Management Plans (DES,
                 2018)

South Australia  Queensland's QCoast2100 program
                 Directions for a Climate Smart South Australia (SA Government, 2019a)

Tasmania         Climate Action 21: Tasmania's Climate Change Action Plan 2017­2021 (State of
                 Tasmania, 2017a)

                 Tasmania's 2016 State Natural Disaster Risk Assessment (White et al., 2016a)

                 Tasmanian Planning Scheme ­ State Planning Provisions 2017, Coastal Inundation
                 Hazard Code and a Coastal Erosion Hazard Code (Government of Tasmania, 2017).

Victoria         In accordance with the Climate Change Act 2017, Victoria has a Climate Change
                 Adaptation Plan 2017-2020 (Victoria State Government DELWP, 2016) including a
                 Monitoring, Evaluation, Reporting and Improvement (MERI) framework for Climate
                 Change Adaptation in Victoria (DELWP, 2018), Victorian Climate Projections (2019) and
                 multiple resources for regions and local government (Victoria DELWP 2020).

                 Heatwaves in Victoria. A 2018 vulnerability assessment of the state to heatwaves using a
                 Damage and Loss Assessment methodology (Natural Capital Economics, 2018)

Western Australia Western Australian Government Adapting to our changing climate
                         2012 (WA Government, 2016)

                 State Planning Policy 2.6 ­ Coastal Planning (SPP2.6)

Regional and local (examples only)

104 have declared Climate Emergencies to leverage climate action as of September 2021 covering 36.6% of the
Australian population (Climate Emergency Declaration, 2020)

Tasmania         2017: Tasmanian Planning Scheme ­ State Planning Provisions. State of Tasmania, 514.
                 (State of Tasmania 2017) (State of Tasmania, 2017b)

South Australia  Regional integrated vulnerability assessments (IVAs) and adaptation plans (SA
                 Government, 2019a)

NSW              Enabling Regional Adaptation (Jacobs et al., 2016)

Do Not Cite, Quote or Distribute  11-88                                                         Total pages: 151
FINAL DRAFT                              Chapter 11  IPCC WGII Sixth Assessment Report

Victoria       Every region and catchment Management Authority in Victoria has an adaptation plan, as
               does virtually every local government. There are also three alliances of multiple local
NSW            governments working on climate change and new initiatives such as the Climate Change
Queensland     Exchange. https://www.parliament.vic.gov.au/967-epc-la/inquiry-into-tackling-climate-
               change-in-victorian-communities

               Coastal Zone Management Plan for Bilgola Beach (Bilgola) and Basin Beach (Mona
               Vale) (Haskoning Australia, 2016)

               Torres Strait Climate Change Strategy (TSRA, 2014); Torres Strait Regional Adaptation
               and Resilience Plan 2016-2021 (TSRA, 2016)

               Climate Risk Management Framework for Queensland Local Government (Erhart et al.,
               2020)

Northern Territory Climate Change Action Plan (2011-2020) (Darwin City Council, 2011)

1

2 Table 11.15b: Examples of New Zealand's adaptation strategies, plans and initiatives by government agencies at the (a)
3 national level, (b) sub-national, and (c) regional or local level. NB These examples have not been assessed for their
4 effectiveness (see Supplementary Material Table SM11.1b)

Jurisdiction   Strategies/Plans/Actions

New Zealand    The New Zealand Government's adaptation policy framework is based on the following
central        legislation: Resource Management Act 1991; Local Government Act 2002; National
Government     Disaster Resilience Strategy 2019 (CDEM, 2019), and the Climate Change Response (Zero
               Carbon Amendment) Act 2002 (CCRA 2002).

               Adaptation preparedness report 2020/21 baseline is the reporting organisation responses
               from the First Information request under the CCRA 2002 (MfE, 2021) to assist the
               monitoring of progress and effectiveness of adaptation, by the Climate Change Commission

Local          The Department of Conservation's Climate Change Adaptation Action plan sets out a long-
Government     term strategy for climate research, monitoring, and action. DOC climate adaptation plan

               In July 2017, a group of 39 Local Government Mayors and Council Chairs (of 78 in total)
               endorsed a 2015 local government declaration calling for urgent responsive leadership and a
               holistic approach on climate change, with the government needing to play a vital enabling
               leadership role (LGNZ, 2017; Schneider et al., 2017).

               Seventeen councils have declared Climate Emergencies to leverage climate action plans as
               of September 2021, covering 75.3% of the New Zealand population.

                       The MFE adaptation preparedness report states that 18% of councils (11 of 61 surveyed in
                     2021) have some sort of plan or strategy to increase resilience to climate impacts (MfE,
                     2021). Out of New Zealand's 15 regional and unitary councils, two councils have climate
                       adaptation strategies in place. One council has conducted a climate risk assessment. and four
                       have one in development. Five councils have climate action plans and three are in
                       development.

Regional Councils (examples only)

Bay of Plenty  Climate Action Plan July 2019 (non-statutory) Climate Action Plan
Regional       Long Term Plan 2018-2028 (LTP)
Council

Waikato
Regional
Council

Do Not Cite, Quote or Distribute         11-89                                         Total pages: 151
   FINAL DRAFT                             Chapter 11                            IPCC WGII Sixth Assessment Report

   Greater           GWRC's Climate Change Strategy (October 2015) Climate change strategy implementation
   Wellington        Hutt River Flood Risk Management Plan
   Regional
   Council

   Unitary Authorities (examples only)

   Auckland          Auckland Unitary Plan
   Council           AUP RPS B10
                     Table B11.9- (bottom of doc)
                     E36. Natural hazards and flooding

   Marlborough Marlborough Environment Plan First to integrate Dynamic Adaptive Pathways Planning
   District Council (DAPP) into Plan policies.

   Gisborne          Tairwhiti Resource Management Plan (District Plan) March 2020

   District Council

   District Council (example only)

   Waimakariri       Infrastructure Strategy in the Long Term Plan 2017.

   District Council Long-Term-Plan-Further-Information-Document-WEB.pdf

                     Page 113/31

1

 2

 3 11.7.2 Barriers and Limits to Adaptation

 4

 5 Major gaps in the adaptation process remain across all sectors and at all levels of decision-making (11.3;
 6 Table 11.115a Table 15b) (very high confidence). Efforts to build, resource and deploy adaptive capacity are
 7 slow compared to escalating impacts and risks (Stephenson et al., 2018; CoA, 2020e). Barriers to effective
 8 adaptation include governance inertia at all levels, hindering the development of careful and comprehensive
 9 adaptation plans and their implementation (Boston and Lawrence, 2018; MfE and Hawke's Bay Regional
10 Council, 2020; White and Lawrence, 2020). Lack of clarity about mandate, roles and leadership, and
11 inadequate funding for adaptation by national and State governments and sectors, are slowing adaptation
12 (Lukasiewicz et al., 2017; Waters and Barnett, 2018; LGNZ, 2019; MfE, 2020c) (11.3; 11.7.1). Established
13 planning tools and measures were designed for static risk profiles, and practitioners are slow to take up tools
14 better suited to changing climate risks (CoA, 2020e; Schneider et al., 2020) (11.5; Box 11.5). The
15 communication of relevant climate change information remains ad hoc (Stevens and O'Connor, 2015;
16 CCATWG, 2017; Palutikof et al., 2019c; Salmon, 2019). In Australia, the lack of national guidance or
17 adaptation laws create barriers to adaptation, reflected in uneven coastal adaptation based on a "wait and
18 see" approach (Dedekorkut-Howes et al., 2020).

19

20

21 Table 11.16: Examples of barriers to adaptation action in the region

   Barrier                                                               Source

   Governments                                                           (Dedekorkut-Howes et al., 2020)

   Lack of consistent policy direction from higher levels and frequent

   policy reversals

   Conflicts between community-based initiatives, City Councils and (Forino et al., 2019)
   business interests

   Different framings of adaptation between local governments (risk) (Smith et al., 2015; Schlosberg et al., 2017;

   and community groups (vulnerability, transformation)                  McClure and Baker, 2018)

   Competing planning objectives                                         (McClure and Baker, 2018)

   Divergent perceptions of risk concepts                                (Button and Harvey, 2015; Mills et al., 2016b;
                                                                         Tonmoy et al., 2018)

   Focus on climate variability rather than climate change               (Dedekorkut-Howes and Vickers, 2017)

   Do Not Cite, Quote or Distribute                         11-90                                   Total pages: 151
FINAL DRAFT                                           Chapter 11            IPCC WGII Sixth Assessment Report

Low prioritization of climate change adaptation among competing (Glavovic and Smith, 2014; Lawrence et al.,

institutional objectives                                             2015; McClure and Baker, 2018)

Constraints in using new knowledge                                   (Temby et al., 2016)

Lack of institutional and professional capabilities and capacity e.g. (Lawrence et al., 2015; Scott and Moloney,

to monitor and evaluate adaptation outcomes                          2021)

Lack of understanding of Indigenous knowledge and practices          (Parsons et al., 2019)

Lack of authority and political legitimacy                           (Hayward, 2008; Boston and Lawrence, 2018;
                                                                     CCATWG, 2018; Parsons et al., 2019)

Fear of litigation                                                   (Tombs et al., 2018; Iorns Magallanes and
                                                                     Watts, 2019; O'Donnell et al., 2019)
The upfront costs of adaptation relative to competing demands on
government expenditure                                               (Gawith et al., 2020; Warren-Myers et al.,
Private sector                                                       2020b)

Governance and policy uncertainty, lack of cross sector              (CCATWG, 2017; Forino et al., 2017; IGCC,
coordination, lack of capital investment in climate solutions        2021a)

Inconsistent hazard information and incomplete understanding of (CCATWG, 2017; Harvey, 2019)
adaptation

Mismatch in duration of insurance cover (annual) lending             (Storey and Noy, 2017; O'Donnell, 2020)
(decades) and infrastructure and housing investment (50-100ys)

Perceived unaffordability of adaptation, lack of client demand and (Hurlimann, 2008; Hurlimann et al., 2018)
awareness of climate change risks and limited and inconsistent
climate risk regulation in the construction industry

Translating information into organisations to address disinterest    (Warren-Myers et al., 2020b; Warren-Myers et
amongst clients in the property industry                             al., 2020a)

Erosion of adaptive capacity and challenges of transformational      (Jakku et al., 2016)
adaptation in agriculture and rural communities

Communities                                                          (Public Participation, 2014; MfE, 2017a; Archie
Nature of government engagement with communities                     et al., 2018; OECD, 2019b)

Lack of clarity regarding roles and responsibilities                 (Gorddard et al., 2016; Elrick-Barr et al., 2017;
Lack of resourcing of adaptation                                     Goode et al., 2017; Waters and Barnett, 2018)

                                                                     (Singh-Peterson et al., 2015; Lukasiewicz et al.,
                                                                     2017; Brookfield and Fitzgerald, 2018)

         Lack of deep engagement with climate change                 (Kench et al., 2018; Pearce, 2018)
         Diverging perceptions, values and goals within communities  (Austin et al., 2018; Fitzgerald et al., 2019;
                                                                     Marshall et al., 2019)
         Inequities within and between communities                   (Eriksen, 2014; Parkinson, 2019)

         Lack of sustained engagement, learning and trust between    (Serrao-Neumann et al., 2020)
         community, scientists and policy makers

1

Do Not Cite, Quote or Distribute                      11-91                                  Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1

 2 There are many barriers to starting adaptation pre-emptively (CCATWG, 2018) (Table 11.16) (very high
 3 confidence). Recent institutional changes in New Zealand indicate that this is changing (11.7.1; Table 15b).
 4 Many groups are yet to engage deeply with climate change adaptation (Kench et al., 2018) and some
 5 adaptation processes are being blocked (Pearce et al., 2018; Garmestani et al., 2019; Alexandra, 2020) or
 6 exploited to deflect from mitigation responsibilities (Smith and Lawrence, 2018; Nyberg and Wright, 2020).
 7 Some actors are resistant to using climate change information (Tangney and Howes, 2016; Alexandra, 2020).
 8 Fear of litigation and demands for compensation can contribute to this reluctance (Tombs et al., 2018;
 9 O'Donnell et al., 2019)and is increasingly inviting litigation and other costs (Hodder, 2019; Bell-James and
10 Collins, 2020). Jurisprudence is evolving from cases on projects, to cases about decision making
11 accountability in the public and private sectors (Bell-James and Collins, 2020; Peel et al., 2020) and rights
12 based cases (Peel and Osofsky, 2018). National and sub-national governments may become exposed to
13 unsustainable fiscal risk as insurers of "last resort", which can lead to inequitable outcomes for vulnerable
14 groups and future generations (11.3.8), path dependencies and negative effects on physical, social, economic
15 and cultural systems (Hamin and Gurran, 2015; Boston and Lawrence, 2018). Cross-scale governance
16 tensions can prevent local adaptation initiatives from performing as intended (Tschakert et al., 2016; Piggott-
17 McKellar et al., 2019). Adaptation that draws on Mori cultural understanding in partnership with local
18 government in New Zealand can lead to more effective and equitable adaptation outcomes (MfE, 2020a).

19

20 Communities' vulnerabilities are dynamic and uneven (high confidence). In Australia, 435,000 people in
21 remote areas face particular challenges (CoA, 2020e). Some groups do not have the time, resources or
22 opportunity to participate in formal adaptation planning as it is currently organised (Victorian Council of
23 Social Service, 2016; Tschakert et al., 2017; Mathew et al., 2018). Linguistically diverse groups can be
24 disadvantaged by social isolation, language barriers, and others' ignorance of the knowledge and skills they
25 can bring to adaptation (Shepherd and van Vuuren, 2014; Dun et al., 2018) (11.1.2). Social, cultural and
26 economic vulnerabilities, biases and injustices, such as those faced by many women (Eriksen, 2014;
27 Parkinson, 2019) and non-heterosexual groups and gender minorities (Dominey-Howes et al., 2016;
28 Gorman-Murray et al., 2017), can deepen impacts and impede adaptation; (Fitzgerald et al., 2019; Marshall
29 et al., 2019) (Cross-Chapter Box GENDER in Chapter 18).

30

31 Potential biophysical limits to adaptation for non-human species and ecosystems where impacts are projected
32 to be irreversible, with limited scope for adaptation, are signalled in key risks 1-4 (11.6). In some human
33 systems, fundamental limits to adaptation include thermal thresholds and safe freshwater (Alston et al.,
34 2018) (Table 11.14) and the inability of some low-lying coastal communities to adapt in-place (Box 11.6)
35 (very high confidence). Some individuals and communities are already reaching their psycho-social
36 adaptation limits (Evans et al., 2016). A lack of robust and timely adaptation means key risks will
37 increasingly manifest as impacts, and numerous systems, communities and institutions are projected to reach
38 limits (Table 11.14, Figure 11.6), compounding current adaptation deficits and undermining society's
39 capacity to adapt to future impacts (very high confidence).

40

41 11.7.3 Adaptation enablers

42

43 Adaptation enablers include understanding relevant knowledge, diverse values and governance, institutions
44 and resources (Gorddard et al., 2016) (very high confidence). Skills and learning, community networks,
45 people-place connections, trust-building, community resources and support and engaged governance build
46 social resilience that support adaptation (Maclean et al., 2014; Eriksen, 2019; Phelps and Kelly, 2019). A
47 multi-faceted focus on the role societal inequalities and environmental degradation play in generating
48 climate change vulnerability can enable fairer adaptation outcomes (McManus et al., 2014; Ambrey et al.,
49 2017; Schlosberg et al., 2017; Graham et al., 2018).

50

51 The feasibility and effectiveness of adaptation options will change over time depending on place, values,
52 cultural appropriateness, social acceptability, ongoing cost-effectiveness, leadership and the ability to
53 implement them through the prevailing governance regime (Singh et al., 2020). The capacity and
54 commitment of the political system can drive early action that can reduce risks (Boston, 2017).

55

56 Decision makers face the challenge of how to adapt when there are ongoing knowledge gaps, and
57 uncertainties about when some climate change impacts will occur and their scale, e.g. coastal flooding (Box

Do Not Cite, Quote or Distribute  11-92       Total pages: 151
    FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 11.6), or extreme rainfall events and their cascading effects (Box 11.4) (very high confidence). No-regrets
 2 decisions are likely to be insufficient (Hallegatte et al., 2012). A perception exists in some sectors that all
 3 climate risks are manageable based on past experience (CCATWG, 2017). Projected impacts, however, are
 4 outside the range experienced, meaning that decisions have to be made now for long-lived assets, land uses
 5 and communities exposed to the key risks (Paulik et al., 2019a; Paulik et al., 2020) often under contested
 6 conditions where adaptation competes with other public expenditure (Kwakkel et al., 2016). New planning
 7 approaches being used across the region, can enable more effective adaptation, e.g. continual iterative
 8 adaptation (Khan et al., 2015) rapid deployment of decision tools appropriate for addressing uncertainties
 9 (Marchau et al., 2019, and transformation of governance and institutional arrangements {Boston, 2018 #444)
10 (Table 11.17). Recognising co-benefits for mitigation and sustainable development can help incentivise
11 adaptation (11.3.5.3, 11.8.2).

12

13

14 Table 11.17: Key enablers for adaptation

    Enabler            Example

    Governance         Clear climate change adaptation mandate
    frameworks         Measures that inform a shift from reactive to anticipatory decision making, e.g. decision tools that
                       have long timeframes
                       Institutional frameworks integrated across all levels of government for better coordination
                       Revised design standards for buildings, infrastructure, landscape such as common land use
                       planning guidance and codes of practice that integrate consideration of climate risks to address
                       existing and future exposures and vulnerability of people, physical and cultural assets
                       (11.3.1, 11.3.2, 11.3.3, 11.3.4.3, 11.3.5, 1.3.6, 11.4.1, 11.4.2, 11.5.1, 11.5.2, 11.6, 11.7.1, 11.7.2,
                       11.8.1, 11.8.2, Table 11.7, Table 11.14, Box 11.1, Box 11.3, Box 11.5, Box 11.6)

    Building capacity  Provision of nationally consistent risk information through agreed methodologies for risk
    for adaptation     assessment that address non-stationarity
                       Targeted research including understanding the projected scope and scale of cascading and
                       compounding risks
                       Education, training and professional development for adaptation under changing risk conditions
                       Accessible adaptation tools and information
                       (11.1.2, 11.3.4, 11.3.5, 11.4.1, 11.5.1, 11.6, 11.7.1, 11.7.2, Table 11.14, Table 11.16, Table 11.18,
                       Box 11.6)

    Community          Community engagement based on principles that consider social and cultural and Indigenous
    partnership and    Peoples' contexts and an understanding of what people value and wish to protect, e.g.
    collaborative      International Association of Public Participation (Public Participation, 2014).
    engagement         Use of collaborative and learning-oriented engagement approaches tailored for the social and
                       informed by the cultural context
                       Community awareness and network building
                       Building on Indigenous Australian and Mori communities' social-cultural networks and
                       conventions that promote collective action and mutual support
                       (11.3.5, 11.4, 11.7.1, 11.7.3.2, Table Box 11.1.1, Table 11.14, Box 11.6)

    Dynamic adaptive Increased understanding and use of decision-making tools to address uncertainties and changing
    decision-making risks, such as scenario planning and dynamic adaptive pathways planning to enable effective

                            adaptation as climate risk profiles worsen
                            (11.7.3.1, 11.7.3.2, Table 11.14, Table 15b, Table 11.18, Box 11.4, Box 11.6)

    Funding            Adaptation funding framework to increase investment in adaptation actions
    mechanisms         New private sector financial instruments to support adaptation
                       (11.7.1, 11.7.2, Table 11.16)

    Reducing           Economic and social policies that reduce income and wealth inequalities
    systemic           Strengthening social capital and cohesion
    vulnerabilities    Identifying and redressing rigid or fragmented administrative and service delivery systems
                       Review of land use and spatial planning to reduce exposure to climate risks
                       Restoring degraded ecosystems and avoiding further environmental degradation and loss.
                       (11.1.1,11.1.2,11.3.5, 11.3.11, 11.4.1, 11.5.1.3, 11.7.2, 11.8.1, Table 11.10, Table 11.13)

15
16

    Do Not Cite, Quote or Distribute  11-93                                                       Total pages: 151
FINAL DRAFT                       Chapter 11                         IPCC WGII Sixth Assessment Report

 1 11.7.3.1 Planning and Tools

 2

 3 Adaptation decision support tools enable a shift from reactive to anticipatory planning for changing climate
 4 risks (high confidence). The available tools are diversifying with futures and systems methodologies and
 5 dynamic adaptive policy pathways being increasingly used (Bosomworth et al., 2017; Prober et al., 2017;
 6 Lawrence et al., 2018a; CoA, 2020e; Rogers et al., 2020a; Schneider et al., 2020) (11.5; Box 11.6) to help
 7 shift from static to dynamic adaptation by highlighting path dependencies and potential lock in of decisions,
 8 system dependencies and the potential for cascading impacts (Table 11.17) (Wilson et al., 2013; Clarvis et
 9 al., 2015; Pearson et al., 2018; Cradock-Henry et al., 2020b; Lawrence et al., 2020b). Modelling and tools to
10 test the robustness and cost-effectiveness of options (Infometrics and PSConsulting, 2015; Qin and Stewart,
11 2020) can be used alongside adaptation strategies with decision-relevant and usable information (Smith et
12 al., 2016; Tangney, 2019; Serrao-Neumann et al., 2020), particularly when supported by effective
13 governance and national and sub-national guidance (Box 11.6).

14

15 More inclusive, collaborative and learning-oriented community engagement processes are fundamental to
16 effective adaptation outcomes (11.7.3.2) (Boston, 2016; Lawrence and Haasnoot, 2017; Sellberg et al., 2018;
17 Serrao-Neumann et al., 2019a; Simon et al., 2020) (very high confidence). More participatory vulnerability
18 and risk assessments can better reflect different knowledge systems, values, perspectives, trade-offs,
19 dilemmas, synergies, costs and risks (Jacobs et al., 2019; Ogier et al., 2020; Tonmoy et al., 2020). A shift
20 from hierarchical to more cooperative governance modalities can assist effective adaptation (Vermeulen et
21 al., 2018; Steffen et al., 2019; CoA, 2020e; Lawrence et al., 2020b; MfE, 2020a; Hanna et al., 2021).

22

23 Regular monitoring, evaluation, communication and coordination of adaptation are essential for accelerating
24 learning and adjusting to dynamic climate impacts and socio-economic and cultural conditions change
25 (Moloney and McClaren, 2018; Palutikof et al., 2019a; Cradock-Henry et al., 2020a) (high confidence).
26 Training to improve decision-makers' `evaluative capacity' can play a role (Scott and Moloney, 2021).
27 Climate action benchmarking, diagnostic tools and networking can enhance the adaptation process across
28 diverse decision settings e.g. water, coasts, protected areas and Indigenous Peoples (Ayre and Nettle, 2017;
29 Davidson and Gleeson, 2018; Coenen et al., 2019; Gibbs, 2020). Effective adaptation requires cross-
30 jurisdictional and cross-sectoral policy coherence and national coordination (Delany-Crowe et al., 2019;
31 Rychetnik et al., 2019; MfE, 2020c).

32

33

34 Table 11.18: Examples of adaptation decision tools

Tools               Application                              Source

Scenario analysis,  For futures planning in coastal, urban,  (Randall et al., 2012; Jones et al., 2013; CSIRO,
modelling, futures  agriculture and health sectors           2014; Bosomworth et al., 2015; Infometrics and
narratives                                                   PSConsulting, 2015; Knight-Lenihan, 2016; Maier
                                                             et al., 2016; Stephens et al., 2017; B. Frame et al.,
                                                             2018; Stephens et al., 2018; Ausseil et al., 2019a;
                                                             Coulter et al., 2019; Serrao-Neumann et al.,
                                                             2019b)

Dynamic Adaptive    For conditions of deep uncertainty for   (Cradock-Henry et al., 2018b; Cradock-Henry et
Pathways Planning   short-term and long-term options and     al., 2020a) (agriculture); (Lawrence et al., 2019b)
(DAPP)              flexibility, and with communities        (flood risk management)
                                                             (Lawrence and Haasnoot, 2017; Colliar and
                                                             Blackett, 2018) (coastal communities)
                                                             (Tasmanian Climate Change Office, 2012; Lin et
                                                             al., 2017; Ramm et al., 2018) (capacity building)
                                                             (Moran et al., 2014; Colloff et al., 2016; Dunlop et
                                                             al., 2016; Bosomworth et al., 2017) (natural
                                                             resource, management)
                                                             (Hadwen et al., 2012; Barnett et al., 2014b; Fazey
                                                             et al., 2015; Lazarow, 2017; Ramm et al., 2018)
                                                             (coastal)
                                                             (Siebentritt et al., 2014; Zografos et al., 2016)
                                                             (regional development)

Do Not Cite, Quote or Distribute  11-94                              Total pages: 151
FINAL DRAFT                                       Chapter 11               IPCC WGII Sixth Assessment Report

                                                              (Maru et al., 2014) (disadvantaged communities)
                                                              (Hertzler et al., 2013; Sanderson et al., 2015)
                                                              (agriculture)
                                                              (Ren et al., 2011) (infrastructure and resilient
                                                              cities)
                                                              (Cunningham et al., 2017) (social network
                                                              analysis with communities)

Serious Games       To catalyse learning, raise awareness and (Lawrence and Haasnoot, 2017; Colliar and

                    explore attitudes and values              Blackett, 2018; Flood et al., 2018; Edwards et al.,

                                                              2019)

Signals and Triggers For where there is near-term certainty   (Stephens et al., 2017; Stephens et al., 2018)
for monitoring DAPP and longer-term deep uncertainty e.g.

                            sea-level rise

Shared Socio-       For where there is deep uncertainty and (B. Frame et al., 2018)

economic Pathways scenarios are used

Hybrid Multi-criteria For conditions of deep uncertainty for  (D. Frame et al., 2018; Lawrence et al., 2019a)
analysis and DAPP short-term and long-term options and
(deep uncertainty) flexibility desired

Real Options        For conditions of deep uncertainty        (Infometrics and PSConsulting, 2015; Infometrics,
Analysis (ROA)                                                2017; Lawrence et al., 2019a; Wreford et al.,
                                                              2020)

Scenario-based cost- For conditions of deep uncertainty       (Guthrie, 2019)
benefit analysis

Portfolio analysis  For uncertainties in the land use sector (Monge et al., 2016; Awatere et al., 2018)(West et
                                                                        al. 2021)

Cost Benefit        Where decisions can be easily reversed    (Hadwen et al., 2012; Little and Lin, 2015;
Analysis                                                      Stewart, 2015; Luo et al., 2017; Thamo et al.,
                    For assessing and prioritising physical   2017)
Vulnerability       and social place-based risks, using       (Ramm et al., 2017; Moglia et al., 2018; Pearce et
assessment          indices, modelling and participatory      al., 2018; Tonmoy and El-Zein, 2018)
                    approaches
Statutory tools     For planning direction                    (DoC NZ, 2010; DoC NZ, 2017a; DoC NZ,
                                                              2017b; NSW Government, 2018)
                    For planning and design of adaptation
                                                              (MfE, 2017a)

Standards           For adaptation best practice              (ISO, 2019)
Jurisprudence
                    For adaptation implementation and legal (O'Donnell and Gates, 2013; McAdam, 2015;
Guidance
                    interpretation                            Iorns Magallanes and Watts, 2019; Peel et al.,

                                                              2020)

                    For adaptation and use of uncertainty     (CSIRO and BOM, 2015; MfE, 2017a; Lawrence
                    tools                                     et al., 2018b; Palutikof et al., 2019b)

Information delivery For adaptation decision making           https://coastadapt.com.au/
and decision support
portal

Do Not Cite, Quote or Distribute                  11-95                                   Total pages: 151
    FINAL DRAFT                              Chapter 11                  IPCC WGII Sixth Assessment Report

    Monitoring,          For local government, private sector and  (Goodhue et al., 2012; Little et al., 2015; IGCC,
    evaluation and       finance sector to benchmark, track        2017; Lawrence et al., 2020a; LGAQ and DES,
    reporting on         progress                                  2020; Rogers et al., 2020b; WAGA, 2020)
    adaptation progress                                            (Moloney and McClaren, 2018)
    (incl. adaptation
    indices and web-
    based tools)

1

2

3 11.7.3.2 Attitudes, Engagement and Accessible Information as Enablers

4

 5 Concern for climate change has become widespread (Hopkins, 2015; Borchers Arriagada et al., 2020), giving
 6 climate adaptation social legitimacy (high confidence). Over three quarters of Australians (77%) agree that
 7 climate change is occurring and 61% believe climate change is caused by humans (Merzian et al., 2019). A
 8 growing proportion of Australians perceive links between climate change and high temperatures experienced
 9 during heatwaves and extremely hot days (2018/2019 Summer) (48%), droughts and flooding (42%), and
10 urban water shortages (30%) (Merzian et al., 2019). Rural populations in NSW perceive climate change
11 impacts as stressing their wellbeing and mental health and requiring leadership and action (Austin et al.,
12 2020). In New Zealand, between 2009 and 2018, the proportion of New Zealanders who agreed or strongly
13 agreed that climate change is real increased from 58% to 78% (a 34.5% increase), while those agreeing or

14 strongly agreeing with human causation increased from 41% to 64% (a 56.1% increase) (Milfont et al.,
15 2021). Nevertheless, New Zealanders have a tendency to overestimate the amount of sea-level rise,
16 especially amongst those most concerned about climate change, and incorrectly associate it with melting sea

17 ice, which has implications for engagement and communication strategies (Priestley et al., 2021).

18

19 Use of more systemic, collaborative and future-oriented engagement approaches is facilitating adaptation in
20 local contexts (Rouse et al., 2013; MfE, 2017a; Leitch et al., 2019) (high confidence). Local `adaptation

21 champions' and experimental and tailored engagement processes can enhance learning (McFadgen and

22 Huitema, 2017; Lindsay et al., 2019). Dynamic adaptive pathways planning (Lawrence et al., 2019a) and
23 inclusive community governance (Schneider et al., 2020)can help progress difficult decisions such as the
24 relocation of cultural assets and managed retreat, and contestation about which public goods to prioritise and
25 how adaptation should be implemented (Kwakkel et al., 2016) (Colliar and Blackett, 2018). Participatory
26 climate change scenario planning can test assumptions about the present and the future (Mitchell et al., 2017;
27 Serrao-Neumann and Choy, 2018; Chambers et al., 2019; Serrao-Neumann et al., 2019c) and help envision
28 people-centred, place-based adaptation (Barnett et al., 2014b; Lindsay et al., 2019). Social network analysis
29 can inform engagement and communication of adaptation (Cunningham et al., 2017). Knowledge brokers,

30 information portals and alliances can help communities, governments and sector groups to better access and

31 use climate change information (Shaw et al., 2013; Fünfgeld, 2015; Lawrence and Haasnoot, 2017). Novel
32 approaches to building climate change literacy and adaptation capability go hand in hand with dedicated
33 expert organisational support (Stevens and O'Connor, 2015; CCATWG, 2018; Palutikof et al., 2019c;

34 Salmon, 2019). All of these approaches depend on adequate resourcing (very high confidence).

35

36 11.7.3.3 Knowledge Gaps and Implementation Enablers

37

38 There are two priority areas where new knowledge is critical for accelerating adaptation implementation.

39

40 1) System complexity and uncertainty in observed and projected impacts

41  · Regionally relevant projections of rainfall, runoff, compound and extreme weather (11.2.1; 11.3.3;

42  Box 11.4).

43  · Inclusion of cascading and compounding impacts in integrated assessments (11.5.1) including for

44  infrastructure (11.3.5), tourism (11.3.7) and health (11.3.6) and for different groups, including

45  Aboriginal and Torres Strait Islander Peoples and Tangata Whenua Mori communities (11.4).

46  · Impacts on terrestrial and freshwater ecosystems, including in-situ monitoring to detect ongoing

47  changes especially in New Zealand (11.3.1), and marine biodiversity including environmental

48  tolerances of key life stages (11.3.2).

49  · Repository of indigenous species distribution data for monitoring responses to climate change and

50  climate advisory services for New Zealand (11.3.1.3).

    Do Not Cite, Quote or Distribute         11-96                         Total pages: 151
    FINAL DRAFT                               Chapter 11           IPCC WGII Sixth Assessment Report

1   · National risk assessment for Australia (11.7.1).

2   · The interactions between adaptation and mitigation, particularly where land carbon mitigation is

3   impacted by climate change (11.3.4.3; Box 11.5).

4

5 2) Supporting adaptation decision making

6   · Better understanding of who and what is exposed and where, and their vulnerability to climate

7   hazards (11.3, 11.4).

8   · National assessments of the costs and benefits of climate change, with and without different levels

9   and timings of adaptation and mitigation (11.5.2.3) (11.7.1).

10  · Understanding available adaptation strategies and options, their feasibility and effectiveness as the

11  climate changes, including their intended and unintended outcomes (11.7, 11.8).

12  · Understanding how to embed robust planning approaches into decision making that retain flexibility

13  to change course in the future (11.7.1).

14  · Mechanisms for sharing knowledge and practice of adaptation (11.7).

15  · The role of development paradigms, values and political economy in adaptation framing and

16  effective implementation (11.8).

17  · Understanding social transitions and social licence, for timely, robust and transformational

18  adaptation (11.8.2).

19

20

21 11.8 Climate Resilient Development Pathways

22

23 Adaptation to climate risks and global mitigation of greenhouse emissions determine whether development
24 pathways are climate resilient (Chapter 18). In the near-term, progress towards climate resilient development
25 can be monitored by progress on the Sustainable Development Goals (SDGs). According to government

26 reports (OECD, 2019a) (Figure 11.6) current and projected trajectories fall short of meeting all targets (Allen
27 et al., 2019). Key climate risks for the region (11.6, Table 11.14) affect all of the SDGs, and pre-existing

28 societal inequalities exacerbate climate risks (11.3.5). Projected climate risks combined with underlying
29 SDG indicators will increasingly impede the region's capacity to achieve and maintain a number of SDGs,
30 including sustainable agriculture, affordable and clean energy, sustainable cities and communities, life below

31 water and life on land (OECD, 2019a). Reducing these risks would require significant and rapid emission
32 reductions to keep global warming to 1.5-2.0oC, and robust and timely adaptation (IPCC, 2018).

33

34 11.8.1 System Adaptations and Transitions

35

36 A step-change in adaptation action is needed to address climate risks and to be consistent with climate

37 resilient development (very high confidence). Current adaptation falls short on assessment of complex risks,
38 implementation, monitoring, and evaluation. It is largely incremental and temporary given the scale of
39 projected impacts, it has limits and is mainly reactive rather than anticipatory. Furthermore, risks are

40 projected to cascade and compound, with impacts and costs that challenge adaptive capacities (11.5) and call
41 for transformational responses (11.6, Table 11.15a; Table 11.15b; Supplementary Tables SM11.1a;
42 SM11.1b).

43

44 Current global emissions reduction policies are projected to lead to a global warming of 2.1-3.9 °C by 2100

45 (Liu and Raftery, 2021), leaving many of the region's human and natural systems at very high risk and
46 beyond adaptation limits (high confidence). With higher levels of warming, adaptation costs increase, loss

47 and damages grow, and governance and institutional responses have reduced adaptive capacity. Underlying

48 social and economic vulnerabilities and injustices further reduce adaptive capacity, exacerbating
49 disadvantage in particular groups in society. Sustainable development across and beyond the region will help

50 reduce shared adaptation challenges (11.5.1.2). Effective adaptation avoids lock-in and path dependency,

51 reduces vulnerabilities, increases flexibility to change, builds adaptive capacity and progresses SDGs, thus
52 improving intra- and inter-generational justice (11.5, 11.6, 11.7). Reducing greenhouse gas emissions and

53 structural inequalities is key to achieving the SDGs and contributing to climate resilient development.

54

55 Integrated and inclusive adaptation decision making can contribute to climate resilient development by better
56 mediating competing values, interests and priorities and helping to reconcile short-and long-term objectives,

57 as well as public and private costs and benefits, in the face of rapidly and continuously changing risk profiles

    Do Not Cite, Quote or Distribute            11-97                                Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

 1 (Gorddard et al., 2016; MfE, 2017a; Schlosberg et al., 2017) (11.5.2) (very high confidence). Use of new
 2 tools and approaches (Table 11.18) to address system interactions that match the scale and scope of the
 3 problem can result in more effective adaptation, including proactive and anticipatory governance and
 4 institutional enablers (11.7, Table 11.17) (Schlosberg et al., 2017; Boston and Lawrence, 2018). Building
 5 cities and settlements that are resilient to the impacts of climate change requires the simultaneous
 6 consideration of infrastructural, ecological, social, economic, institutional, and political dimensions of
 7 resilience including political will, leadership, commitment, community support, multilevel governance, and
 8 policy continuity (Torabi et al., 2021).

 9

10 11.8.2 Challenges for Climate Resilient Development Pathways

11

12 Implementing enablers can help drive adaptation ambition and action consistent with climate resilient
13 development (11.7.3, Table 11.17) (very high confidence). However, the scale and scope of cascading,
14 compounding and aggregate impacts (11.5.1) calls for new and timely adaptation, including more effective
15 ongoing monitoring, evaluation, review and continual adjustment (11.7.3) towards the transformations that
16 can break through the `path dependencies' that define the way things are done now (Cradock-Henry et al.,
17 2018b; UN et al., 2018; Head, 2020). However, complex interactions between objectives can create social
18 and economic trade-offs (Table 11.1, 11.3.5.3, 11.7.3.1, Box 11.6).

19

20 Delay in implementing climate change adaptation and emissions reductions will impede climate resilient
21 development, resulting in more costly climate impacts and greater scale of adjustments in the future (IPCC,
22 2018) (11.5.1; 11.5.2) (Box 11.6) and legal risks for those with adaptation mandates and for financial
23 institutions (11.5.1) (very high confidence). The scale and scope of societal change needed for the region to
24 transition to more climate resilient development pathways requires close attention to governance, ethical
25 questions, the role of civil society, the place of Aboriginal and Torres Strait Islander Peoples and Tangata
26 Whenua Mori in the co-production of ongoing adaptation at multiple scales (Koehler et al., 2017; Loorbach
27 et al., 2017; Hill et al., 2020).

28

29 The region faces an extremely challenging future that will be highly disruptive for many human and natural
30 systems (IPCC, 2018) (UNEP, 2020; AAS, 2021; IPCC, 2021) (11.5.1; 11.6; 11.7) (Box 11.1-11.6) (Table
31 11.14). The extent to which the limits to adaptation are reached depends on whether global warming peaks
32 this century at 1.5, 2 or 3+°C above pre-industrial levels. Whatever the outcome, adaptation and mitigation
33 are essential and urgent. (very high confidence)

34

35

36 [START FAQ11.1 HERE]

37

38 FAQ 11.1: How is climate change affecting Australia and New Zealand?

39

40 Climate change is affecting Australia and New Zealand significantly. Some natural systems of cultural,
41 environmental, social and economic significance are at risk of irreversible change. The socio-economic
42 costs of climate change are substantial, with impacts that cascade and compound across sectors and
43 regions, as demonstrated by heatwaves, wildfire, cyclone, drought and flood events.

44

45 Temperature has increased by 1.4°C in Australia and 1.1°C in New Zealand over the last 110 years, with
46 more extreme hot days. The oceans in the region have warmed significantly, resulting in longer and more
47 frequent marine heatwaves. Sea levels have risen and the oceans have become more acidic. Snow depths
48 have declined and glaciers have receded. North-western Australia and most of southern New Zealand have
49 become wetter, while southern Australia and most of northern New Zealand have become drier. The
50 frequency, severity and duration of extreme wildfire weather conditions has increased in southern and
51 eastern Australia and north-eastern New Zealand.

52

53 The impacts of climate change on marine, terrestrial and freshwater ecosystems and species are evident. The
54 mass mortality of corals throughout the Great Barrier Reef during marine heatwaves in 2016­2020 is a
55 striking example. Climate change has contributed to the unprecedented south-eastern Australia wildfires in
56 the spring-summer of 2019­2020, loss of alpine habitats in Australia, extensive loss of kelp forests, shifts
57 further south in the distribution of almost 200 marine species, decline and extinction in some vertebrate

Do Not Cite, Quote or Distribute  11-98       Total pages: 151
FINAL DRAFT                                          Chapter 11  IPCC WGII Sixth Assessment Report

 1 species in the Australian wet tropics, expansion of invasive plants, animals and pathogens in New Zealand,
 2 erosion and flooding of coastal habitats in New Zealand, river flow decline in southern Australia, increased
 3 stress in rural communities, insurance losses for floods in New Zealand, increase in heat wave mortalities in
 4 Australian capital cities, and the fish deaths in Murray-Darling River in the summer of 2018­2019.

 5

 6 [END FAQ11.1 HERE]

 7

 8

 9 [START FAQ11.2 HERE]

10

11 FAQ 11.2: What systems in Australia and New Zealand are most at risk from ongoing climate change?

12

13 The nine key risks to human systems and ecosystems in Australia and New Zealand from ongoing climate
14 change are shown in Figure FAQ 11.2.1. Some risks, especially on ecosystems, are now difficult to avoid.
15 Other risks can be reduced by adaptation, if global mitigation is effective.

16

17 Risk is the combination of hazard, exposure and vulnerability. For a given hazard (e.g. fire), the risk will be
18 greater in areas with high exposure (e.g. many houses) and/or high vulnerability (e.g. remote communities
19 with limited escape routes). The severity and type of climate risk varies geographically (Figure FAQ 11.2.1).
20 Everyone will be affected by climate change, with disadvantaged and remote people and communities the
21 most vulnerable.

22

23 The risks to natural and human systems are often compounded by impacts across multiple spatial and
24 temporal scales. For example, fires damage property, farms, forests and nature with short- and long-term
25 effects on biodiversity, natural resources, human health, communities and the economy. Major impacts
26 across multiple sectors can disrupt supply chains to industries and communities and constrain delivery of
27 health, energy, water and food services. These impacts create challenges for adaptation and governance of
28 climate risks. When combined, these have far-reaching socio-economic and environmental impacts.

29

30

31

32 Figure FAQ 11.2.1: Key risks from climate change

33

34

35 [END FAQ11.2 HERE]

36

Do Not Cite, Quote or Distribute                     11-99       Total pages: 151
    FINAL DRAFT                       Chapter 11               IPCC WGII Sixth Assessment Report

1

2 [START FAQ11.3 HERE]

3

4 FAQ 11.3: How can Indigenous Peoples' knowledge and practice help us understand contemporary

5   climate impacts and inform adaptation in Australia and New Zealand?

6

7 In Australia and New Zealand, as with many places around the world, Indigenous Peoples with connections

8 to their traditional country and extensive histories, hold deep knowledge from observing and living in a

9 changing climate. This provides insights that inform adaptation to climate change.

10

11 Indigenous Australians - Aboriginal and Torres Strait Islanders - maintain knowledge regarding previous
12 sea-level rise, climate patterns, and shifts in seasonal change associated with flowering of trees and
13 emergence of food sources, developed over thousands of generations of observation of their traditional
14 country. Knowledge of localised contemporary adaptation is also held by many Indigenous Australians with
15 connections to traditional lands. With assured Free and Prior-Informed Consent, this provides a means for
16 Indigenous-guided land management, including for fire management and carbon abatement, fauna studies,
17 medicinal plant products, threatened species recovery, water management, and weed management.

18

19 Tangata Whenua Mori in New Zealand are grounded in Mtauranga Mori knowledge which is based on
20 human-nature relationships and ecological integrity and incorporates practices used to detect and anticipate

21 changes taking place in the environment. Social-cultural networks and conventions that promote collective

22 action and mutual support are central features of many Mori communities and these customary approaches
23 are critical to responding to, and recovering from, adverse environmental conditions. Intergenerational
24 approaches to planning for the future are also intrinsic to Mori social-cultural organisation and are expected

25 to become increasingly important, elevating political discussions about conceptions of rationality, diversity

26 and the rights of non-human entities in climate change policy and adaptation.

27

28

29 [END FAQ11.3 HERE]

30

31

32 [START FAQ11.4 HERE]

33

34 FAQ 11.4: How can Australia and New Zealand adapt to climate change?

35

36 There is already work underway by governments, businesses, communities and Indigenous Peoples to help
37 us adapt to climate change. However, much more adaptation is needed for the ongoing and intensifying

38 climate risks. This includes coordinated laws, plans, guidance and funding that enable society to adapt, and

39 the information, education and training that can support it. Everyone has a role to play, working together.

40

41 We currently mainly react to climate events such as wildfires, heatwaves, floods and droughts, and generally

42 rebuild in the same places. However, climate change is making these events more frequent and intense, and

43 ongoing sea-level rise and changes in natural ecosystems are advancing. Better coordination and
44 collaboration between government agencies, communities, Aboriginal and Torres Strait Islanders and
45 Tangata Whenua Indigenous Peoples, not-for-profit organisations and businesses will help prepare for these

46 climate impacts more proactively, in combination with future climate risks integrated into their decisions and
47 planning. This will reduce the impacts we experience now and the risks that will affect future generations.

48

49 Some of the risks for natural systems are close to critical thresholds and adaptation may be unable to prevent
50 ecosystem collapse. Other risks will be severe, but we can reduce their impact by acting now, for example

51 coastal flooding from sea-level rise, heat-related mortality and managing water stresses. Many of the risks
52 have potential to cascade across social and economic sectors with widespread societal impacts. In such cases,
53 really significant system-wide changes will be needed to the way we live and govern currently. To facilitate
54 such change, new governance frameworks, nationally consistent and accessible information, collaborative

55 engagement and partnerships with all sectors, communities and Indigenous Peoples and the resources to

56 address the risks, are needed (Figure FAQ 11.4.1).

57

    Do Not Cite, Quote or Distribute                   11-100                         Total pages: 151
FINAL DRAFT                       Chapter 11  IPCC WGII Sixth Assessment Report

1 However, our ability to adapt to climate change impacts also rests on every region in the world playing its
2 part in reducing greenhouse gas emissions. If mitigation is ineffective, global warming will be rapid,
3 adaptation costs will increase, with worsening losses and damages.

4
5

 6
 7

 8 Figure FAQ 11.4.1: Developing adaptation plans in the solutions space showing system tipping points, thresholds and
 9 limits to adaptation, unsustainable pathways, critical systems and enablers to climate resilient development

10

11

12 [END FAQ11.4 HERE]

13

14

Do Not Cite, Quote or Distribute  11-101      Total pages: 151