What is the most frequent and widespread disaster causing hazard?

2.1 Australia has a long history of disasters that are linked to natural hazards.

2.2 Natural disasters are more than just natural hazards. Disaster risk is a product of the type and intensity of the natural hazard event, the extent to which communities and other assets are exposed to a natural hazard event, and the vulnerability or ability of communities and other systems to cope with and recover from the impacts of the natural hazard event.

2.3 The extent of damage and harm caused by natural disasters depends on a wide range of factors, such as the intensity of the disaster, where people choose to live, how they build their homes, how both public and private land is managed, and how well people and communities are prepared, supported and cared for during and after disasters.

2.4 Climate change has already increased the frequency and intensity of extreme weather and climate systems that influence natural hazards.

2.5 Further global warming over the next two decades is inevitable. As a result, sea‑levels are projected to continue to rise. Tropical cyclones are projected to decrease in number, but increase in intensity. Floods and bushfires are expected to become more frequent and intense.

2.6 The 2019‑2020 bushfire season demonstrated that bushfire behaviour is becoming more extreme and less predictable. Catastrophic fire conditions may become more common, rendering traditional bushfire prediction models and firefighting techniques less effective.

2.7 We can also expect more concurrent and consecutive hazard events. For example, in the last 12 months there was drought, heatwaves and bushfires, followed by severe storms, flooding and a pandemic. Concurrent and consecutive hazard events increase the pressure on exposed and vulnerable communities. Each subsequent hazard event can add to the scale of the damage caused by a previous hazard event. There are likely to be natural disasters that are national in scale and consequence.

2.8 As 2020 has shown, some communities will have to cope with the effects of multiple nature hazard events at once, with the prospect of being affected by further hazard events before the recovery efforts have been completed.

2.9 To properly manage natural disasters of national scale and consequence, it is no longer suitable or appropriate to assess disaster risk at an individual hazard level. We must assess the risk of multiple hazard events occurring concurrently or consecutively. We must look for opportunities to reduce the exposure of communities to natural hazard events and increase the capacity of communities to prepare for and recover from their impacts.

Australia is prone to natural hazards

2.10 Australia’s variable climate, geography and environment place Australian communities, infrastructure, ecosystems and cultural and heritage values in the path of frequent and high-energy natural hazard events.

2.11 Natural hazards are driven primarily by weather and geology. Examples of weather‑driven natural hazards include bushfire, flood, heatwave, cyclones, landslides, east coast lows and thunderstorms. Geological-driven hazards include earthquakes and tsunami.

The Australian experience

2.12 Australia has a naturally variable climate, [6] with temperature and rainfall fluctuating from season to season and year to year. It is common for large parts of the country to move from hot, dry conditions (heatwaves and droughts) to cool, wet conditions (often associated with floods). Natural hazards, including bushfires, floods and heatwaves, are linked to these drought and flood cycles.

2.13 While the focus of our inquiry has been the devastating 2019-2020 bushfire season, Australia has no shortage of significant weather events. For example, since the commencement of our inquiry in late February 2020 Australia has experienced the following significant weather events:

  • In March 2020, Victoria, Queensland and WA felt the effects of ex‑tropical cyclone Esther, experiencing daily rainfall records, flooding and building damage. Melbourne had its wettest March day since 1929, affecting infrastructure services. [7]
  • In April 2020, NSW, Victoria and the ACT were confronted by strong cold fronts, wind and rain, with some sites experiencing their coldest April days since the 1950s. Tasmania also had stronger than average winds, with gusts up to 146km/h. [8]
  • In May 2020, across the country rainfall was, on the whole, below average, but the effects of this were concentrated in the southern parts of the country. Most of the north experienced above average rainfall. [9]
  • In June 2020, across the country, but particularly in WA and the north, it was much warmer than average. WA was buffeted by storms and damaging winds in the latter half of the month, causing property damage across the state and loss of power to over 3,000 properties. [10]
  • In July 2020, two low pressure systems in the Tasman Sea brought heavy rain, isolated flash flooding and high seas along the NSW coast, and snow in its alpine regions. An extensive dust storm swept across Australia’s interior, with visibility severely reduced – many locations reporting less than 200 metres. [11]
  • In August 2020, a strong cold front produced severe wind gusts along the west coast of WA, with wind gusts raising dust, bringing down trees and causing roof damage. Snow settled in Launceston, Tasmania, with the most comparable event likely from 1921. A few stations in WA and the NT had their highest August mean daily maximum temperature on record. [12]

2.14 In September 2020, the Bureau of Meteorology identified the return of La Niña and forecast above average rainfall over eastern and northern Australia in the coming spring and summer. We also heard evidence that the return of La Niña has seen a ‘very active’ hurricane season for the United States and Mexico and ‘multiple typhoons’ through East Asia, which may relate to ‘an early onset and more active tropical cyclone season’ for Australia. [13]

2.15 Australia also experiences concurrent and consecutive natural hazard events. Concurrent and consecutive hazard events increase the pressure on exposed and vulnerable communities. Each subsequent hazard event can add to the scale of the damage caused by the previous hazard event. This was the case over the 2019‑2020 summer, during which communities experienced successive conditions of drought, heatwaves, bushfires, hailstorms, and flooding; compounding the destructive impact on the affected communities.

2.16 Geological seismic events also frequently occur in Australia. Approximately 100 earthquakes, with a magnitude of three or more, occur in Australia each year. [14] Australia’s largest recorded earthquake was in 1988 at Tennant Creek in the NT, with an estimated magnitude 6.6, but it occurred in a sparsely populated area. [15] By contrast, the Newcastle earthquake in 1989 had a magnitude of 5.4, claimed 13 lives and caused widespread damage to building and facilities. [16]

2.17 Each state and territory varies in its experience of natural hazards. Each has its own climate, geography and environment that influence the type, frequency, intensity and severity of hazards experienced. As a result, the resourcing and arrangements in mitigating and responding to natural hazards also varies in each jurisdiction.

Natural hazards have already increased and intensified

2.18 We heard from the Commonwealth Scientific and Industrial Research Organisation (CSIRO), Australia’s national science research agency, that climate change is adding to Australia’s natural climate variability, driving changes in average and extreme weather, and increasing climate impacts on our water resources, ecosystems, health, infrastructure and economy, both now and continuing into the future. [17]

2.19 Clear trends have emerged in recent decades beyond the ‘noise’ of Australia’s natural climate variability. [18] Warming is an ongoing trend – Australia has warmed by approximately 1.4 degrees since 1910. [19] As shown in Figure 2, 2019 was Australia’s hottest year on record. [20]

Figure 2: Australian mean temperature anomaly, 1910-2019 [21]

2.20 There is also a drying trend across much of the southern half of Australia, particularly in the south west and particularly in the winter months. [22] Over the last 20 years, the southern half of Australia experienced below average rainfall. [23] This is the most sustained large-scale change in rainfall since national records began in 1900. [24]

2.21 These trends are influencing fire seasons, heatwaves, rainfall and flood risk. [25] Observed changes that are being influenced by background climate trends include:

  • increased frequency of heatwaves and record high temperatures
  • longer fire seasons with more extreme fire danger days
  • increase in marine heatwaves, and
  • reduced annual average rainfall in some regions. [26]

2.22 Events that are starting to be influenced by climate trends include:

  • an increase in heavy rainfall, and
  • increased frequency of coastal storm surge inundation. [27]

2.23 In some regions, there may also be a trend towards more weather-dominated fire events. In weather-dominated events, fires interact with the atmosphere resulting in unpredictable and extreme fire behaviour. [28] The most extreme of these are known as firestorms or pyrocumulonimbus (pyroCb) events, [29] which can be associated with extraordinarily destructive fire behaviour. Figure 3 shows the occurrence of pyroCb events in Australia, including the highest annual record of pyroCb events in 2019.

Figure 3: Register of Australian pyrocumulonimbus wildfire (pyroCb) events: 1994‑2019 [30]

2.24 Professor Jason Sharples, University of NSW, said that the number of these thunderstorms, which form in the smoke plume of a fire, in the 2019‑2020 bushfires was ‘unprecedented’. [31] We heard evidence of ‘multiple’ pyroCb events during the 2019‑2020 bushfires. [32] Professor David Bowman, University of Tasmania, told us:

… this last summer there was a near doubling of the record of these events in one event, and that assembly of data goes back about 30 years. So something happened this last summer which is truly extraordinary because - what we would call statistically a black swan event - we saw a flock of black swans. That just shouldn’t have happened. [33]

Climate-driven natural hazards are expected to become more frequent and intense

2.25 We heard that Australia’s climate is ‘virtually certain’ to get warmer. [34] Ongoing drying of the climate of much of southern and eastern Australia is likely. Other threats include ongoing sea-level rise and an increase in extreme weather events such as short-duration heavy rainfall. [35]

2.26 To assess climate risk, governments, businesses and others rely on multiple lines of evidence, including climate models that provide projections of the likely changes in the climate in the future. Climate models are continuously being updated and developed as different modelling groups around the world incorporate new computing technologies, higher spatial resolution, and new information on physical and biogeochemical processes.

2.27 There are significant projected changes from the present for many hazards. Figure 4 shows the projected changes for Australia’s climate-driven natural hazards over coming decades. It is drawn from the climate scenario analyses released in September 2020 as part of the Climate Measurement Standards Initiative, an industry-led collaboration between insurers, banks, scientists, regulators, reporting standard professionals, service providers and supporting parties. The analyses assessed and synthesised multiple lines of evidence and the existing scientific literature. We discuss the Climate Measurement Standards Initiative further in Chapter 4: Supporting better decisions.

Figure 4: Projected changes in climate-driven natural hazards. Confidence estimates are provided in brackets. [36]

2.28 There are three factors that determine the climate, and hence climate risks, in future scenarios: ongoing natural climate variability; global socio-economic development and the resulting emissions of greenhouse gases and aerosols; and how the climate responds at a regional level to these emissions. [37]

2.29 The 2018 State of the Climate Report, issued by the CSIRO and the Bureau of Meteorology (BoM) stated that the ‘amount of climate change expected in the next decade or so is similar under all plausible global emissions pathways’. [38] Globally, temperatures will continue to increase, and Australia will have more hot days and fewer cool days. [39]

2.30 Climate projections, such as those from CSIRO and the Bureau of Meteorology often draw on an ‘ensemble’ (or collection) of models from around the world called the Coupled Model Intercomparison Project (CMIP), which includes the Australian model ACCESS. CMIP5 is the current global ensemble. The future climate information in the 2018 State of the Climate Report used CMIP5 modelling. [40] CMIP6 is part of updated climate modelling, intended to improve on CMIP5, and is being undertaken to inform the 2021 Intergovernmental Panel on Climate Change sixth assessment report (AR6). [41]

Figure 5: Climate projections for Australia [42]

2.31 Figure 5 shows the projected change in average temperature in Australia using CMIP5 and CMIP6. It shows the projections using CMIP5 and CMIP6 across ‘low’ (green) and ‘high’ (red) greenhouse gas emission scenarios.

2.32 An initial assessment of the CMIP6 is that it improves the confidence of the current CMIP5 projections and may provide for greater precision at a regional level, and many results from both sets of modelling are consistent with each other. [43]

2.33 The climate response under both emission pathways are very similar in the next 20 years but then diverge after that. [44] We heard from the BoM that further ‘warming over the next two decades is inevitable’[45] and that over the next 20 to 30 years, ‘the global climate system is going to continue to warm in response to greenhouse gases that are already in the atmosphere’. [46] We heard from CSIRO that some further climate change is ‘locked in’, ‘because of emissions we’ve already had’. [47]

2.34 We heard from CSIRO that even under the low emissions scenario, which goes to net negative emissions, the climate does not return to a preindustrial or recent baseline type climate immediately. It takes a very long time for that to occur, and would require CO2 to be removed from the atmosphere. [48] According to CSIRO, it is ‘more a matter of stabilising rather than returning’. Australia ‘need[s] to adapt to further changes in the climate no matter what happens with emissions and we will have inevitable changes in the climate coming through for decades to come, no matter what pathway we take forward’. [49]

2.35 Strong adaptation measures are necessary to respond to the impacts of climate change.

2.36 Warming beyond the next 20 to 30 years is largely dependent on the trajectory of greenhouse gas emissions. [50]

Outlooks for different natural hazards

Individual natural hazard outlooks

2.37 The outlook for changes in frequency and intensity of hazards varies. The sections below summarise outlooks for common natural hazards in Australia based on current information.

Tropical cyclones

2.38 Climate models project a future decrease in the total number of tropical cyclones, but an increase in their intensity. [51] However, large natural variability and data limitations make it difficult to be entirely confident about long-term trends in tropical cyclones. [52] Despite this, coastal impacts from tropical cyclones are likely to become worse, due to rising sea levels and increases in cyclone-related extreme rain and wind events. [53]

2.39 Additionally, the latitude at which tropical cyclones reach their maximum lifetime intensity may be shifting poleward (in Australia, towards the south), and a further movement poleward is possible in future. This could have serious consequences for south-eastern Queensland and north-eastern NSW which are reasonably densely populated areas. [54] In the west coast, south of Shark Bay, the consequences of a poleward shift are likely to be smaller, yet still significant for the central west and lower west coasts of WA due to the interactions of tropical lows and mid-latitude weather systems. [55]

Storms and rainfall events

2.40 Extreme rainfall events are projected to increase in intensity, potentially resulting in an increase in flood risks. [56]

2.41 Already, there is evidence that the proportion of total annual rainfall coming from heavy rain days has increased. [57]

2.42 In some areas in southern Australia, particularly south west WA, the increased risk of extreme rainfall events may be partly offset by the projected decrease in average rainfall. [58]

Coastal flooding and inundation

2.43 By 2090, the Australian sea level is projected to rise by between 26 and 82 cm depending on the level of emissions and how relevant systems respond. [59] A greater sea level rise is possible if ice sheets melt faster than projected. [60]

2.44 The consequences of sea level rise for Australia will include the flooding of low lying coastal and tidal areas with increased regularity. [61] It is also likely to result in coastal erosion, loss of beaches and higher storm surges that will affect coastal communities, infrastructure, industries and the environment. [62]

Earthquakes and tsunami

2.45 Australia’s earthquake prone regions extend from the southern highlands down to Gippsland in Victoria, the Mount Lofty and Flinders Ranges in SA, the WA wheat belt and north-west of WA. [63]

2.46 Earthquakes, as well as landslides and volcanic activity, can cause tsunami – waves that are generated when the ocean is disturbed over a broad area in a short period.

2.47 In 2018, Geoscience Australia conducted a National Seismic Hazard Assessment, with contributions from experts on seismology. It provides an improved understanding of earthquakes in Australia. Under the National Seismic Hazard Assessment, it is forecast that over a 50 year period, on average, 10% of Australia is likely to experience earthquakes exceeding the expected peak intensity for a given location, though the magnitude of these earthquakes may still be relatively low. [64]

Heatwaves

2.48 Heatwaves are commonly defined as three or more days of consistently high temperatures that are unusual for a region. Other heat indices include overnight minimum temperatures, and other factors relating to human comfort (eg humidity). Heatwaves are Australia’s deadliest natural disaster, accounting for almost five times more fatalities than bushfires. [65]

2.49 Heatwave events have increased in intensity, frequency and duration across Australia in recent decades. [66] Hot temperatures are occurring earlier in spring, and later in autumn. 2019 was Australia’s hottest year on record, with a record 42 days when Australia’s area-averaged daily mean temperature was above the 99th percentile. [67]

2.50 Further warming over the next two decades is inevitable, in response to past and future greenhouse gas emissions. Hot days, warm spells and heatwaves are all projected to occur more often and with increased intensity. Extreme hot days that now occur every 20 years are expected to occur every two to five years by 2050. [68]

Bushfires

2.51 Fire weather is primarily a function of temperatures, humidity and winds. [69] There has been a long-term increase in dangerous fire weather, and in the length of the fire season, across large parts of Australia. [70] There has been a reduction in the time between the catastrophic bushfire events of Australian history. [71]

2.52 In Australia, changes in fire danger risk are often assessed using trends in the Forest Fire Danger Index (FFDI), which uses temperature, humidity, wind speed and rainfall to assess fire danger. In southern and eastern Australia, the length of the fire season, as measured using the FFDI, has increased in recent decades. [72] The fire weather season now arrives more than three months earlier than in the mid‑twentieth century in some parts of Australia. [73] The lengthening of the fire season is reducing the opportunities to undertake prescribed burning, and this is likely to get worse in the future.

2.53 There has been an increase in the frequency and severity of fire weather since 1950 in southern and eastern Australia, and this trend is projected to continue. [74]

Figure 6: Future projections of fire weather conditions show increasing fire danger days [75]

GCMs = Global Climate Models. CCAM = conformal Cubic Atmospheric Model (CSIRO regional climate modelling).

NARCLiM = NSW and ACT regional climate modelling project.

2.54 Fire danger is very likely to increase in the future for many regions of Australia. [76] The increased frequency of days with a high FFDI is likely to result in reduced intervals between fire events, and increase fire intensities, which could make fighting fires harder. [77]

2.55 In northern and central Australia, monsoonal rainfall and the spread of invasive weeds (such as gamba grass) has increased in recent decades, resulting in increased fuel growth. This may also lead to more dangerous fire conditions during the dry season. [78]

2.56 Projections for changes in fire conditions in northern and central Australia are less certain than for southern and eastern Australia, as the incidence of fire is strongly related to fuel availability and the occurrence of episodic rain events, however, predictions indicate that dry season fires will be more dangerous. [79]

2.57 Climate projections show that more dangerous weather conditions for bushfires are very likely to occur throughout Australia in the future due to a warming climate. The change in climate is also likely to result in changes to the amount, structure and type of bushfire fuel. [80] Climate models also indicate a future increase in dangerous pyro-convection conditions for many regions of southern Australia. [81]

2.58 Pyro-convection can lead to the formation of pyroCb clouds and fire‑generated thunderstorms, as shown in Figure 7. Compared to previous fire seasons, they were relatively common across the 2019‑2020 bushfire season. [82] Fire‑generated thunderstorms can cause a rapid and dangerous escalation of fire, which becomes highly dynamic and unpredictable in its behaviour. They can increase fire spread by lofting embers and creating destructive wind gusts and tornadic vortices. These fire-generated thunderstorms also generate lightning, which can start new fires. [83] For example, during the Black Saturday fires (Victoria, 2009), lightning generated in the fire plume ignited a new fire about 100kms ahead of the main area of the fire front. [84]

Figure 7: Formation of pyrocumulonimbus clouds and fire‑generated thunderstorms [85]

When natural hazards become disasters

2.59 Natural hazards on their own are not disasters – they are merely earth systems in operation. Disaster occurs when natural hazards intersect with people and things of value, and when impacts of hazards exceed our ability to avoid, cope or recover from them.

2.60 In 2015, Australia and other members of the United Nations adopted the Sendai Framework for Disaster Risk Reduction 2015 – 2030 (the Sendai Framework). Through the Sendai Framework, countries around the world recognise the importance of managing all aspects of disaster risk:

Over the last twenty years thinking about how to reduce disaster losses has greatly expanded beyond a simple focus on disaster management to consideration of all the other elements that contribute to increasing the risk of loss of life, injury, damage to critical infrastructure and economic losses when disaster strikes. [86]

2.61 As we have heard, there are three factors that contribute to disaster risk:

  • natural hazards – a natural process or phenomenon that may cause loss of life, injury or other adverse impacts, including on mental and physical health, property, the economy, communities, and environmental assets[87]
  • exposure – people, property or other assets present in hazard areas that are subject to potential losses, and
  • vulnerability – the conditions determined by physical, social, economic and environmental factors or processes which increase the susceptibility of an individual, community, assets or systems to the impacts of hazards. [88]

2.62 Disaster risk can, therefore, be managed by focusing efforts toward each of these factors. These risk factors can be managed at each of the ‘before’, ‘during’ and ‘after’ stages of natural disasters: see Figure 8.

2.63 The recovery phase from one disaster can also be the mitigation and preparedness phases of the next.

Figure 8: Elements of disaster risk associated with natural hazards [89]

2.64 In 2018, a National Disaster Risk Reduction Framework (NDRRF) was developed by the Australian Government with representatives from all levels of government, business and the community sector. In March 2020, the Prime Minister and first ministers of the Australian, state and territory governments endorsed the NDRRF.

2.65 The NDRRF states that all sectors of society must work together to reduce disaster risk, and sets out approaches for doing so across four priority areas: understanding disaster risk; ensuring accountable decisions; enhancing investment; and providing governance, ownership and responsibility. The NDRRF concludes that, although a shared responsibility, disaster risk is often not shared equally. It notes that institutional decision-making often places the risk on communities and individuals, who have varying capacity to manage it:

While individuals and communities have their roles to play, they do not control many of the levers needed to reduce some disaster risks. Governments and industry in particular must take coordinated action to reduce disaster risks within their control to limit adverse impacts on communities. [90]

2.66 As disaster risk increases, the capacity of communities and systems to be resilient is diminished.

Disaster impacts are extensive, complex and long-term

2.67 Disasters involving natural hazards mark Australian history – through lives taken; physical and psychological injuries caused; homes destroyed; animals killed and injured; ecosystems damaged; and heritage and cultural sites damaged or destroyed.

Table 1: Fatalities in Australia between 1900 and 2015 resulting from natural hazards [91]

Natural hazard

Fatalities (1900-2015)

Extreme heat

4,555

Flood

1,911

Tropical cyclone

1,216

Bushfire

974

Lightning

562

Gust

527

Landslide

96

Tornado

53

Earthquake

17

Hail

3

2.68 In late 2017, Deloitte Access Economics estimated that, for the preceding decade, natural disasters have cost Australia $18.2 billion per year on average, taking into account both tangible and intangible costs. [92]

2.69 Figure 9 shows insured losses across the decade from 2010-2020. We recognise that there are broader direct and indirect costs from these events, but nonetheless insured losses usefully illustrate the significant toll disasters have taken in Australia.

Figure 9: Insured losses from natural hazards – 2010 to 2020 [93]

2.70 Fatalities and economic loss are common measures of impact. However, the full impacts of natural disasters are difficult to capture in quantifiable metrics.

2.71 The 2019‑2020 bushfires illustrate the significant challenges posed to individuals, communities, businesses and governments to withstand and rebuild from natural disasters. We heard that the impacts of disasters can be long-term, complex and intangible. See Chapter 21: Coordinating relief and recovery.

2.72 Australia’s recent and still unfolding history is a useful illustration of how resilience can be stretched or exceeded due to consecutive events and compounding impacts. Australian individuals, communities and businesses have been impacted by fire, flood, drought and a global pandemic within the last 12 months – and, for many, the impacts have been concurrent or consecutive.

Australia’s disaster outlook is alarming

2.73 CSIRO’s recent Climate and Disaster Resilience Technical Report to the Prime Minister concluded:

The COVID-19 pandemic is a timely reminder of how hazards within the complex and changing global risk landscape can affect lives, livelihoods and health. It provides a compelling case for an all-hazards approach to achieve risk reduction as a basis for sustainable development. [94]

Climate and disaster risks are growing across Australia. This is due to intensifying natural hazards under a changing climate and increasing exposure and vulnerability of people, assets, and socio-economic activities in expanding hazard areas. [95]

2.74 Direct and indirect disaster costs in Australia are projected to increase from an average of $18.2 billion per year to $39 billion per year by 2050, even without accounting for climate change. [96] The costs associated with natural disasters include significant, and often long-term, social impacts, including death and injury and impacts on employment, education, community networks, health and wellbeing.

2.75 A recent analysis by Risk Frontiers, Macquarie University and the Bushfire and Natural Hazards Cooperative Research Centre (BNHCRC) examined Australia’s history of ‘compound disasters’. It identified disasters occurring with other societal stressors such as wars, recessions and pandemics further exacerbating their consequences. The analysis defined these as a ‘compound disaster’, the ‘combining of numerous drivers and/or hazards that add to societal or environmental risk’ (see Figure 10). [97]

Figure 10: Compound disasters from 1966-1967 to 2018-2019 within a three-month window, and at least ten normalised deaths and/or $100 million normalised losses. [98]

2.76 The analysis concluded:

The occurrence of compound disasters at the time of societal stressors would further amplify impacts and result in complex emergency management challenges. Consideration of the coincidence of other societal stressors at the time of compound disasters has not received attention before the current COVID-19 pandemic. [99]

2.77 It is no longer suitable to assess disaster risk at an individual hazard level, without taking account other possible natural hazards and broader societal pressures that impact resilience.

2.78 As hazards become more frequent and intense, measures to prevent or mitigate hazards will become more difficult. [100]

2.79 For example, floods can be mitigated by levees and dams; the intensity and rate of spread of bushfires can, in some but not all circumstances, be reduced through fuel management; and the severity of heatwaves in residential areas can be reduced by good urban design. For other natural hazards such as storms, cyclones and earthquakes, our ability to influence them is limited or non-existent.

Exposure

2.80 Climate and hazard risk analysts at Cross Dependency Initiative assessed disaster risk across 15 million addresses in 544 local government areas, looking across 2020 and 2100, analysing data for five hazards across Australia. The analysis indicates that there are a large number of current and projected assets exposed to natural hazards:

There are 383,300 addresses in 2020 which would be classified as High Risk Properties. This number is projected to increase to 735,654 in 2100 for existing development only. This figure does not account for new development occurring in high hazard areas, or continued use of inadequate building standards, which unabated will substantially increase this number. [101]

2.81 Land-use planning is the primary mechanism that governments can use to manage exposure to natural hazards. Land-use planning governs how land can be used, where built assets can be located, and how they are designed.

2.82 Land-use planning decisions have far-reaching and long-lasting consequences as to how exposed and vulnerable the community will be to future natural hazards. Where land-use planning decisions do not effectively incorporate natural hazard risk, future impacts of natural disasters will be higher. We discuss land-use planning decisions further in Chapter 19: Land-use planning and building regulation.

Vulnerability and resilience

2.83 Vulnerability can be physical and relate to the susceptibility to damage of the built environment. This includes the vulnerability of infrastructure systems where damage to components disrupt service delivery. Vulnerability also includes the vulnerability of people and the likelihood of injury or death in a natural hazard event. [102]

2.84 Vulnerability is closely related to the concept of resilience, which is the ability of a system, community or society exposed to hazards to resist, absorb, accommodate, adapt to, transform and recover from the effects of a hazard in a timely and efficient manner, including through the preservation and restoration of its essential basic structures and functions through risk management. [103]

2.85 For the most part, the lifestyles and daily activities of Australians are heavily dependent on interconnected systems for the delivery of essential services
(eg energy, water, food, health and education services, transport, and communications). [104] These systems support communities and influence their vulnerability and resilience to disasters. [105]

2.86 The National Resilience Taskforce noted that the changing nature of many hazards, coupled with growing and ageing populations and infrastructure in exposed areas, is leading to increased exposure and vulnerability. [106]

2.87 In July 2020, the University of New England (UNE) and the BNHCRC launched the National Disaster Resilience Index. The index measures resilience through a combination of social, economic, natural environment, built environment, governance and geographical factors (for example, access to information and availability of emergency services).

2.88 The index assesses disaster resilience according to a set of coping and adaptive capacities:

Coping capacity is the means by which available resources and abilities can be used to face adverse consequences that could lead to a disaster. Adaptive capacity is the arrangements and processes that enable adjustment through learning, adaptation and transformation. [107]

2.89 The work by UNE and the BNHCRC found that there is a general pattern of higher capacity for disaster resilience across the populated south east areas of Australia, and around metropolitan and major regional centres (Figure 11). Their research findings included:

  • most of the population of Australia live in areas assessed as having ‘moderate’ capacity for disaster resilience
  • a ‘low’ capacity for disaster resilience is associated with remote and very remote areas, comprising a total of about 435,000 people
  • areas with ‘low’ disaster resilience comprise over 93% of Australia’s land surface area, and
  • almost 50% of non-metro areas have a ‘low’ capacity for resilience; less than 10% of metro areas have a ‘low’ capacity for resilience.

2.90 Disaster resilience is a complex interplay of factors, including social and economic characteristics, the provision of government and other services, community capital and governance regimes.

2.91 Managing disaster risk and resilience will require a greater focus on managing the factors that contribute to them and, in particular, those factors over which we have some control.

Figure 11: Australia’s capacity for disaster resilience [108]

Toplist

Latest post

TAGs