4613.0 - Australia's Environment: Issues and Trends, Jan 2010  
ARCHIVED ISSUE Released at 11:30 AM (CANBERRA TIME) 28/01/2010   
   Page tools: Print Print Page Print all pages in this productPrint All

This document was added or updated on 05/02/2010.

Image: Earth FEATURE ARTICLE: CLIMATE CHANGE IN AUSTRALIA


ISSUE: CLIMATE CHANGE

CLIMATE CHANGE IN AUSTRALIA
What is climate change?
Australia's greenhouse gas emissions
Reducing our greenhouse gas emissions
Impacts and adaptation
Summary
Endnotes

ISSUE: CLIMATE CHANGE

Climate change is a global problem with global consequences. Warmer-than-average temperatures are being recorded across the world. Glaciers and polar ice caps are melting and sea levels are rising. Mounting evidence indicates that these changes are not the result of the natural variability of climate.

The International Panel on Climate Change (IPCC), established in 1988 by the World Meteorological Organisation (WMO) and the United Nations Environment Program (UNEP), released its fourth assessment report in 2007. It declared that ‘warming of the climate system is unequivocal’ and it is ‘very likely’ that greenhouse gas emissions from human activity have caused most of the observed global temperature increase since the mid-1900s (Endnote 1).

In Australia and internationally, there has been an increasing focus on the issue of climate change and the demand for credible statistics and information has grown accordingly.

This feature article begins with a brief discussion of the science of climate change, followed by a statistical examination of Australia’s contribution to global greenhouse gas emissions, and opportunities for reducing emissions in Australia. The last section presents statistics related to the impacts climate change is projected to have on Australia’s society, economy and environment and some broad adaptation measures being undertaken.
  • What is climate change? While climate changes can occur naturally, there is now general agreement that global warming over the last 50 years is very likely the result of human activities, specifically the emission of greenhouse gases into the atmosphere. Increased levels of greenhouse gases in the atmosphere trap heat and increase the earth’s temperature. Since 1950, Australia’s average annual temperature has increased by 0.9°C. If global emissions remain high, by 2070 the average annual temperature is projected to increase by a further 2.2 to 5.0°C (Endnote 2).
  • Australia’s greenhouse gas emissions: Australia has about 0.3% of the world’s population, but contributes about 1.5% of total greenhouse gas emissions (Endnote 3). This puts Australians among the highest per capita emitters. In 2007, Australia’s net greenhouse gas emissions across all sectors totalled 597.2 million tonnes of carbon dioxide equivalent (Mt CO2-e) under the accounting provisions of the Kyoto Protocol (Endnote 4).
  • Reducing our greenhouse gas emissions: Reducing greenhouse gas emissions is necessary to mitigate human-induced climate change. There are many opportunities for households and businesses in Australia to reduce emissions, including large-scale use of renewable energy sources, improving energy efficiency and greater use of public transport. Atmospheric levels of greenhouse gases can also be potentially reduced by activities which increase the amount of carbon stored in our soils and forests. Putting a price on carbon emissions would change the relative prices of different forms of energy and accelerate the move to a low carbon economy.
  • Impacts and adaptation: Australia’s climate is already changing and in coming decades the Australian community will probably need to take steps to adapt to the impacts of climate change that cannot be avoided by mitigation. Some of the areas considered most vulnerable to the impacts of climate change include water, agriculture, biodiversity, coastal settlements and human health. In some cases, households and businesses are already taking voluntary action to adapt to a changing climate. Some areas are vulnerable precisely because their capacity to adapt is limited.


CLIMATE CHANGE IN AUSTRALIA

WHAT IS CLIMATE CHANGE?

The Intergovernmental Panel on Climate Change (IPCC) defines climate change as a change in the state of the climate that can be identified by changes in the mean and/or the variability of its properties, and that persists for an extended period, typically decades or longer” (Endnote 5). Climate change is also often referred to as global warming.

Globally, there is evidence of climate change, including increases in air and ocean temperatures, widespread melting of snow and ice and rising sea levels (Endnote 6).

The Greenhouse Effect

The earth’s atmosphere is like a blanket that keeps the planet warm. Incoming energy from the sun penetrates the atmosphere to warm the earth. The earth then radiates heat back toward space. Some of the outgoing heat is absorbed by greenhouse gases in the atmosphere and re-emitted back to earth, keeping the planet at a level warm enough to support life. This is called the greenhouse effect.

An enhanced greenhouse effect can cause climate change as increased levels of greenhouse gases (mostly carbon dioxide) in the atmosphere lead to more heat being trapped, so the earth’s temperature increases. While natural phenomena have caused the climate to change in the past, there is now a general consensus that human activities are largely responsible for today’s very high levels of greenhouse gas emissions and associated climate change.


IMPORTANT TERMS
Adaptation – adjustments in natural or human systems in response to actual or anticipated climate changes or their effects.
Carbon sink – a natural or human activity or mechanism that removes carbon dioxide from the atmosphere, such as the absorption of carbon dioxide by growing trees.
Climate in a narrow sense is usually defined as the ‘average weather’ a region experiences, usually calculated over a 30-year period. It usually encompasses surface variables such as temperature, precipitation and wind. While weather can vary dramatically from one day to the next, climate cannot.
Extreme weather event – meteorological conditions which are rare for a particular place and/or time, such as an intense storm or heat wave. An extreme climate event is an unusual average over time of a number of weather events, for example heavy rainfall over a season.
Greenhouse gases – both natural and anthropogenic gases in the atmosphere that absorb and emit infrared or heat radiation, causing the greenhouse effect. The main greenhouse gases are water vapour, carbon dioxide (CO2), methane (CH4) and nitrous oxide (N2O).
Mitigation – refers to response strategies that aim to limit human-induced climate change by reducing greenhouse gas emissions or removing greenhouse gases from the atmosphere through sequestration.
Sequestration refers to the uptake and storage of carbon. For example, trees and plants absorb carbon dioxide, release the oxygen and store the carbon in above-ground organic matter or in the soil. In the context of response strategies, sequestration usually refers to the process of increasing the storage of carbon (for example via reforestation), increasing the carbon content of the soil, or removal of carbon dioxide from flue gases for storage below ground or in the deep ocean.
Weather is the state of the atmosphere at a given time and place. It refers to the temperature, air pressure, humidity, wind, cloudiness and precipitation of a region over a short period of time (e.g. daily maximum temperature).
Note: Many of these definitions are in a climate change context and may not apply in other fields.
Source: Australian Greenhouse Office, 2003, Climate Change: An Australian Guide to the Science and Potential Impacts.
Australia’s changing climate

“Global warming is real, humans are very likely to be causing it, and …it is very likely that there will be changes in the global climate system in the centuries to come larger than those seen in the recent past”.
CSIRO and Bureau of Meteorology, 2007 (Endnote 7).

Australia’s climate is changing. Since 1950, Australia’s average annual temperature has increased by 0.9°C. Scientists conclude that most of this change is likely due to human emissions of greenhouse gases. Declines in snow cover, increases in warm days and decreases in cold days are also likely to be attributable to human activity (Endnote 8).

Rainfall is also changing but the causes of these changes can be difficult to determine. Studies have estimated that 50% of the decrease in rainfall in south-western WA over the last 30 years has been due to human-induced climate change (Endnote 9). Recent CSIRO climate modelling indicates effects on Australia’s climate due to aerosol pollution (Endnote 10) from the northern hemisphere. These include increased rainfall in north-western Australia, and increased air pressure over southern Australia, leading to less rainfall there (Endnote 11).

ANNUAL MEAN TEMPERATURE ANOMALIES 1910 TO 2008

Graph: Annual mean temperature anomalies 1910 to 2008
Note: Anomalies are based on 1961 to 1990 average of 21.8ºC.
Source: Bureau of Meteorology, 2009, Australian Climate Change and Variability, <http://reg.bom.gov.au/silo/products/cli_chg>, last viewed October 2009.


Predicting future climate change

Climate models are used by scientists to simulate the climate system and predict how greenhouse gas emissions are likely to change the climate in the future (Endnote 12).

Building on global scientific knowledge, CSIRO and the Australian Bureau of Meteorology have used climate models to project future climate change in Australia. Their 2007 report presents climate change projections to 2070 for a range of emissions scenarios (Endnote 13):
  • assuming a low emissions scenario, by 2050 annual warming is projected to increase by 0.8 to 1.8°C and by 2070 to 1.0 to 2.5°C;
  • assuming a high emissions scenario, annual warming is projected to increase by 1.5 to 2.8°C by 2050 and by 2.2 to 5.0°C by 2070.

Australia is a vast continent and the projected changes to climate vary considerably from region to region and from season to season. For example, over the next few decades warming is expected to be greater in inland areas than in coastal areas and rainfall is projected to change little in the far north but decrease elsewhere (Endnote 14).

Other projected changes – the magnitude of which depends on the emissions scenario – include:
  • increases in the frequency of hot days and warm nights;
  • changing rainfall, (e.g. by 2070, under the high emissions scenario, rainfall in southern areas is projected to change by between -30% and +5%);
  • decreases in snow cover, average snow season lengths and peak snow depths;
  • increases in annual potential evapotranspiration (the transport of water into the atmosphere from the earth’s surfaces and vegetation);
  • increases in the occurrence of drought, particularly in south-western Australia;
  • increased fire weather risk in some areas; and
  • global seal level rise of 18 to 59 cm, with an additional contribution of up to 17 cm from ice sheet dynamics (Endnote 15). However, larger values cannot be excluded (Endnote 16), and recent research indicates that a sea level rise of one metre or more by 2100 is possible (Endnote 17).


AUSTRALIA’S GREENHOUSE GAS EMISSIONS

Greenhouse gases are produced by human activities such as burning of fossil fuels (e.g. coal, oil and gas), agriculture and land clearing. The concentration of greenhouse gases in the atmosphere varies naturally over time, but since around 1750, human activities associated with industrialisation have dramatically increased these concentrations. For example, concentrations of carbon dioxide now far exceed the natural range over the last 650,000 years (Endnote 18).

While Australia only accounts for around 1.5% of global greenhouse gas emissions, its per capita (per person) CO2 emissions are nearly twice the OECD average and more than four times the world average (Endnote 19).

Australia’s relatively high per capita emissions can be attributed to factors such as the high usage of coal in electricity generation and agricultural emissions from large numbers of sheep and cattle (Endnote 20).

Australia’s Department of Climate Change provides annual estimates of Australia’s greenhouse gas emissions, under the accounting rules that apply for the Kyoto Protocol (Endnote 21).

Australia’s net greenhouse gas emissions in 2007 totalled 597.2 Mt (million tonnes) of CO2-e (carbon dioxide equivalent). This represented a 9% increase from the 1990 level of 547.7 Mt CO2-e (Endnote 22).

AUSTRALIA’S NET GREENHOUSE GAS EMISSIONS 1990 TO 2007

Graph: Australia's net greenhouse gas emissions 1990 to 2007
Note: Kyoto-based estimates of Australia’s net greenhouse gas emissions.
Source: Department of Climate Change, 2009, National Greenhouse Gas Inventory May 2009.

GLOBAL WARMING POTENTIAL OF MAJOR GREENHOUSE GASES
Carbon dioxide (CO2) is the most commonly emitted and probably the best-known greenhouse gas, but there are many others, such as water vapour, methane (CH4), nitrous oxide (N2O), sulphur hexafluoride (SF6), perfluorocarbons (CF4 and C2F6), and hydrofluorocarbons (HFCs).

How much a given mass of a particular greenhouse gas contributes to global warming varies with the type of gas, and so the Global Warming Potential (GWP) index has been developed to place all gases on a common measurement footing. Calculating this index for different gases allows the relative contributions of all greenhouse gases to be expressed in terms of their CO2 equivalence. For example, methane has 21 times the global warming potential (GWP) of CO2. Some substances, such as sulphur hexafluoride, have GWPs thousands of times that of CO2 and are of concern even though they are emitted in small quantities.


Gas100-year global warming potential (GWP)
Carbon dioxide (CO2)1
Methane (CH4)21
Nitrous oxide (N2O)310
Sulphur hexafluoride (SF6)23,900
CF4 6,500
C2F6 9.200
Hydrofluorocarbon (HFC)-2311,700
HFC-1252,800
HFC-134a1,300
HFC-143a3,800
Note: These are the greenhouse gases regulated under the Kyoto Protocol. Solvent and other product use can also cause emissions of NMVOCs (non-methane volatile organic compounds).
Source: Department of Climate Change, 2009, National Inventory Report 2007 Volume 1.

CARBON DIOXIDE EQUIVALENT (CO2-e) EMISSIONS, NET, PER CAPITA AND PER $ GDP
Graph: Carbon dioxide equivalent (CO2-e) emissions, net, per capita and per $ GDP
Note: Graph refers to Kyoto Protocol-based estimates of net greenhouse gas emissions. Index displays emissions as a percentage of emissions in 1990.
(a) GDP used is a chain volume measure; reference year 2007–08.
Source: Australian Bureau of Statistics (ABS), 2008, Australian Historical Population Statistics (cat. no. 3105.0.65.001);
ABS, 2009, Australian National Accounts: National Income, Expenditure and Product September 2009 (cat. no. 5206.0);
Department of Climate Change, 2009, National Greenhouse Gas Inventory May 2009.


Changes in emissions

Although Australia’s annual greenhouse gas emissions increased slightly between 1990 and 2007, the country’s per capita emissions rate decreased by 12%. Despite this reduction, Australia continues to emit a large volume of greenhouse gases per capita, in comparison to other OECD countries (Endnote 23).

The greenhouse gas emissions intensity of the Australian economy, expressed as emissions per dollar of GDP (chain volume measure), declined by 38% over the period 1990 to 2007, from 830 g of carbon dioxide equivalent emissions (CO2-e) per dollar GDP in 1990 to 510 g per dollar GDP in 2007 (Endnote 24). The falling trend in emissions per unit of GDP reflects (Endnote 25):
  • specific emissions management activities across sectors;
  • a decline in net land use, land use change and forestry (LULUCF) emissions over the period; and
  • stronger growth in the services sector of the Australian economy, relative to the more energy-intensive manufacturing sector.

By ratifying the Kyoto Protocol in 2007, Australia agreed to stabilise its emissions (for the five-year commitment period of 2008 to 2012) at no more than 108% of its 1990 (base year) emissions level (Endnote 26).

The October 2008 estimate of Australia’s 1990 net emissions was used to calculate Australia’s target emissions under the Kyoto Protocol. The 1990 emissions were estimated at 547.7 Mt CO2-e, so the target emissions under the Protocol were set at 591.5 Mt CO2-e per year (over the period 2008 to 2012) (Endnote 27).

Australia’s net greenhouse gas emissions between 1990 and 2007 increased by 9%. Therefore, in order to meet its Kyoto target for the 2008 to 2012 period, Australia will need to lower its emissions slightly from the 2007 level.

NET GREENHOUSE GAS EMISSIONS BY SECTOR

Emissions Mt CO2-e
Percent of total emissions
Percent change in emissions
1990
2007
2007
1990 to 2007

Energy
286.4
408.2
68.4
42.5
Stationary energy
195.1
291.7
48.8
49.5
Transport
62.1
78.8
13.2
26.9
Fugitive emissions
29.2
37.7
6.3
29.1
Agriculture
86.8
88.1
14.8
1.5
Land use, land use change and forestry (a)
131.5
56.0
9.4
-57.4
Industrial processes (b)
24.1
30.3
5.1
25.7
Waste
18.8
14.6
2.4
-22.3
Australia's net emissions (c)
547.7
597.2
100.0
9.0

(a) Kyoto Protocol-based figures
(b) Includes confidential N2O emissions from industrial processes and solvent and other product use reported as CO2-e.
(c) Strictly speaking the net credits from land use change and forestry should only enter the account during the first commitment period (2008 to 2012). Their inclusion in this table helps our understanding of Australia’s emissions in relation to the Kyoto emissions target which is 591.5 Mt CO2-e each year over the first commitment period.
Source: Department of Climate Change, 2009, National Greenhouse Gas Inventory May 2009.


Emissions by sector

The Department of Climate Change classifies greenhouse gas emissions (and removals, e.g. by forests acting as carbon sinks) into six sectors. The sectors listed below represent the main human activities contributing to the release or capture of greenhouse gases into or from the atmosphere (Endnote 28):
  • Energy
  • Agriculture
  • Land use, land use change and forestry
  • Industrial processes
  • Waste
  • Solvent and other product use

Energy

The energy sector is responsible for the majority of Australia’s greenhouse gas emissions. In 2007, the production and consumption of energy accounted for 68.4% (408.2 Mt CO2-e) of Australia’s net emissions. Of this, 370.5 Mt of emissions were from the combustion of fossil fuels (principally for electricity generation, transport and manufacturing) and 37.7 Mt were from fugitive emissions (related mainly to coal mining) (Endnote 29). Between 1990 and 2007, energy emissions increased by 42.5%.

Stationary energy

Estimated emissions from stationary energy (i.e. emissions from fuel consumption for electricity generation, fuels consumed in the manufacturing, construction and commercial sectors and other sources like domestic heating) totalled 291.7 Mt CO2-e in 2007, or 48.8% of net national emissions. Emissions from stationary energy increased by 49.5% between 1990 and 2007.

In 2007, electricity generation accounted for 199.5 Mt CO2-e (68.4% of stationary energy emissions, and 33% of Australia’s net emissions) (Endnote 30). While stationary energy emissions increased by 49.5% between 1990 and 2007, electricity generation emissions increased by an even larger percentage: 54% (Endnote 31).

FUELS USED IN AUSTRALIAN ELECTRICITY GENERATION, 2006-07

Source
PJ
Share %

Thermal
Black coal
1,379
56.4
Brown coal
671
27.4
Oil
25
1.0
Gas
284
11.6
Total thermal
2,360
96.4
Renewables
Hydro
52
2.1
Wind (a)
23
0.9
Biomass
5
0.2
Biogas
7
0.3
Total renewables
87
3.6

Note: Figures are for energy input, not output.
(a) Includes solar photovoltaic electricity generation.
Source: Australian Bureau of Agricultural and Resource Economics, 2009, Energy in Australia 2009.

Transport

Transport activity is the other major source of emissions related to the combustion of fossil fuels. Transport contributed 78.8 Mt CO2-e or 13% of Australia’s net emissions in 2007. Emissions from this sector were 26.9% higher in 2007 than in 1990.

Road transport was the main source of transport emissions in 2007, accounting for 68.5 Mt CO2-e or 11.5% of national emissions. Passenger cars were the largest transport source, contributing 41.9 Mt CO2-e (Endnote 32).
Agriculture

The agriculture sector produces most of Australia’s methane and nitrous oxide emissions. Agriculture produced an estimated 88.1 Mt CO2-e emissions or 14.8% of net national emissions in 2007.

AGRICULTURE SECTOR EMISSIONS, 2007

Emissions (Mt CO2-e)
CH4
N20
Total

Enteric fermentation
57.6
-
57.6
Manure management
1.9
1.6
3.5
Rice cultivation
0.2
-
0.2
Agricultural soils
-
15.0
15.0
Prescribed burning of savannas
8.1
3.5
11.6
Field burning of agricultural residues
0.2
0.1
0.3
Total agriculture sector
68.0
20.2
88.1

Source: Department of Climate Change, 2009, National Inventory Report 2007 Volume 1.


Land use, land use change and forestry

The Department of Climate Change prepares submissions of Australia’s greenhouse gas emissions in two ways:
  • according to the rules of the Kyoto Protocol; and
  • according to the guidelines of the United Nations Framework Convention on Climate Change (UNFCCC).

The two methods differ only in the treatment of the land use, land use change and forestry (LULUCF) sector (Endnote 33).

Under UNFCCC methodology, all emissions from the human use of land, and from natural events, are accounted for. However, under Article 3.3 of the Kyoto Protocol, emissions reported from the LULUCF sector are limited to (Endnote 34):
  • afforestation and reforestation (i.e. new forest plantings, which correspond to a negative emissions value); and
  • deliberate deforestation of land that was forest on the 1st of January 1990.

LULUCF EMISSIONS

Graph: LULUCF emissions
Note: Estimates of Australia’s net greenhouse gas emissions from the LULUCF sector, by the UNFCCC reporting method and the Kyoto Protocol method.
Source: Department of Climate Change, 2009, National Greenhouse Gas Inventory May 2009;
Department of Climate Change, 2009, National Inventory Report 2007 Volume 2.


Australia’s net LULUCF emissions (and hence its total net greenhouse gas emissions) are much more variable from year to year under the UNFCCC reporting method than under the Kyoto accounting method. This is because Kyoto-based LULUCF emissions reporting does not include emissions from natural events, such as fire, drought and pest attack, nor does it include emissions from land under continued land use, e.g. the growth, harvesting and regrowth of forests (Endnote 35).

Therefore, Australia’s net emissions in 2007 totalled 825.9 Mt CO2-e under the UNFCCC reporting method but only 597.2 Mt CO2-e under the Kyoto accounting method. The Kyoto-based LULUCF net emissions figure for 2007 was 56.0 Mt CO2-e, whereas the UNFCCC-based total was 284.7 Mt CO2-e (Endnote 36).

Industrial processes

Most greenhouse gas emissions from industrial processes are by-products of production from non-energy related sources. For example, high temperature processing of calcium carbonate to produce quicklime releases carbon dioxide emissions.

Emissions from the industrial processes sector were 30.3 Mt CO2-e in 2007, which was equivalent to 5.1% of net national emissions. This emissions level was 6.2 Mt (26%) higher than in 1990 (Endnote 37). The increase can be attributed to a rise in emissions from the consumption of halocarbons and sulphur hexafluoride (SF6), which are used particularly in refrigeration, air conditioning, foam blowing and aerosols (Endnote 38).

Waste

Waste emissions are predominantly methane and account for less than 3% of Australia’s total emissions. Total waste emissions were 14.6 Mt CO2-e in 2007. Waste emissions have decreased by 4.2 Mt CO2-e (22.5%) since 1990 (Endnote 39).

Waste emissions can be from disposal of solid waste, wastewater handling or waste incineration. Solid waste degrades very slowly and methane emissions continue long after the waste is placed in landfill. For this reason, waste emissions estimates for any year include a significant component of emissions resulting from waste disposal over the previous 50 years. Hence, any change to waste management practices are not likely to affect reported waste emission levels for a number of years (Endnote 40).
REDUCING OUR GREENHOUSE GAS EMISSIONS

“Without effective mitigation, the mainstream science tells us that the impacts of climate change on Australia are likely to be severe”.
Garnaut, 2008 (Endnote 41).

It is difficult to know with certainty to what level greenhouse gas emissions should be reduced. This is partly because of the uncertainty associated with predicting future climate change and partly because views about what constitutes acceptable climate change differ depending on ethical, economic and political judgements (Endnote 42).

The Garnaut Climate Change Review was commissioned in April 2007 to examine the impacts of climate change on the Australian economy and to recommend medium to long-term policies and policy frameworks to improve the prospects of sustainable prosperity (Endnote 43). This included recommendations in relation to two global mitigation goals. One was a target for stabilisation of greenhouse gas concentrations in the atmosphere at 450 parts per million (ppm) of CO2-e; the other a less ambitious target of 550 ppm. These targets are associated with long-term temperature increases in the order of two and three degrees Celsius, respectively (Endnote 44).

Opportunities for reducing greenhouse gas emissions

The Garnaut Review outlined various domestic policy options for reducing Australia’s emissions. Market-based approaches included the introduction of an emissions trading scheme which would establish the right to emit greenhouse gases as a tradeable commodity. Examples of regulatory options included mandatory renewable energy targets and energy efficiency standards for buildings and appliances (Endnote 45).

The following section presents statistics on energy intensity of industry, carbon sequestration, passenger transport, energy sources and efficiency, and putting a price on carbon.


WORLD'S RESPONSE TO CLIMATE CHANGE
1988 – United Nations establishes IPCC
The World Meteorological Organisation and the United Nations Environment Programme (UNEP) established the Intergovernmental Panel on Climate Change (IPCC). The Panel produces periodic assessment reports on scientific information relevant to human-induced climate change.

1992 – Global targets for reducing emissions
At the United Nations Conference on Environment and Development (Earth Summit) held in Rio de Janeiro, the United Nations Framework Convention on Climate Change (UNFCCC) was signed by 154 nations (by December 2007, it had been ratified by 192 countries). It provides the overall policy framework for addressing climate change. The Convention, as originally framed, set no mandatory limits on greenhouse gas emissions for individual nations and contained no enforcement provisions; it is therefore considered legally non-binding.

1997 – Kyoto: Legally binding cuts in emissions
The Kyoto Protocol is linked to the UNFCCC. It sets legally binding commitments for the reduction of four specific greenhouse gases (carbon dioxide, methane, nitrous oxide and sulphur hexafluoride) and two groups of gases (hydrofluorocarbons and perfluorocarbons) for 37 industrialised nations and the European community, as well as general commitments for all member countries. The Kyoto Protocol was adopted in Kyoto, Japan, on 11 December 1997, entered into force on 16 February 2005 and has been ratified by 184 Parties of the UNFCCC Convention.

2006 – The Stern report
The Stern report was published in the UK. It examined the economic impact of climate change and found that the costs of inaction far outweighed the costs of action.

2007 – Bali Roadmap
At the United Nations Climate Change Conference, held in Bali, a decision was made to step up international efforts to combat climate change and lay down measures and obligations for the world, after the first commitment period of the Kyoto Protocol expires at the end of 2012.

2008 – Australia ratifies Kyoto Protocol
Australia's ratification of the Kyoto Protocol came into effect on 11 March 2008. Under the Protocol, Australia has agreed to an annual emissions target of 108% of its 1990 emissions during the 2008 to 2012 period.

2008 – The Garnaut report
The Garnaut Climate Change Review examined the impacts of climate change on the Australian economy and the costs of adaptation and mitigation. It analysed the elements of an appropriate international policy response, and the challenges facing Australia in playing its part in that response.

2009 – Copenhagen
A series of UNFCCC meetings took place throughout 2009. These culminated in the United Nations Climate Change Conference (COP 15) in Copenhagen from 7-18 December. The Conference of the Parties (COP) is an association of all the countries which are Parties to the Convention and is the Convention’s highest decision-making authority.

Energy intensity in Australian industry

The energy intensity of an industry is a measure of the energy consumed to produce one unit of economic output (Endnote 46). Reducing energy intensity would contribute to reducing Australia’s greenhouse gas emissions.

Australia’s energy intensity fell 36% over the 30 years to 2006–07, from 4,880 to 3,100 gigajoules of energy consumed per million dollars of Industry Gross Value Added (Endnote 47).

While most industries’ energy intensity levels fell over the three decades to 2006–07, mining and agriculture increased. Transport and construction experienced large reductions in energy intensity (49% and 74% respectively), while other services fell only 13%. Falls in energy intensity may be attributable to factors other than more efficient use of energy. For example, within the economy it may indicate a shift from manufacturing activities to services or, within an industry such as manufacturing, a shift from heavy to light manufacturing (Endnote 48).

ENERGY INTENSITY, ALL INDUSTRIES, 1976-77 TO 2006-07
Graph: Energy intensity, all industries, 1976-77 to 2006-07
Note: Year refers to financial years, e.g. 1977 refers to 1976–77 financial year.
Source: ABS, 2009, Energy Account 2006–07 (cat. no. 4604.0).
Carbon sequestration and offset opportunities in agriculture and forestry

Increasing the amount of carbon stored, or sequestered, in our soils can reduce the amount of greenhouse gases in the atmosphere and improve agricultural productivity.

Fifty-four per cent of Australia’s land area is used for agriculture (Endnote 49). Management practices, such as increasing perennial vegetation in pastures and maintaining crop residues, can increase soil carbon in agricultural soils. However, the potential for agricultural soils to sequester carbon varies depending on climate, soil type and previous management practices (Endnote 50).

Afforestation provides another way in which carbon can be sequestered. The Australian Bureau of Agricultural and Resource Economics has estimated that between six and 28 million hectares of agricultural land (approximately 1.5% to 7% of all agricultural land) will be economically suitable for afforestation between 2007 and 2050, depending on the price of carbon. These forests (a combination of timber plantations and environmental plantings) would sequester between 296 and 885 Mt of carbon by 2050 (Endnote 51).

The Indigenous community in western Arnhem Land has already taken advantage of opportunities to reduce emissions. The West Arnhem Land Fire Abatement Project undertakes strategic fire management to reduce wildfires over an area of 28,000 square kilometres. These activities reduce greenhouse gas emissions by 100,000 tonnes of CO2-e each year, offsetting some of the emissions from the liquefied natural gas (LNG) plant in Darwin. In return, Darwin LNG is paying the Indigenous fire managers around $1 million a year and bringing new jobs and educational opportunities to the region (Endnote 52).


EMISSIONS AND RURAL LAND USE
Opportunities for reducing emissions
Land clearingReduce or cease land clearing.
Enteric emissions from livestockReduce emissions from ruminant livestock by either use of anti-methanogen technology or shift some meat production from sheep and cattle to kangaroos.
Savanna burningChange management to reduce emissions from savanna burning.
BiofuelsReplace fossil fuels with biodiesel made from algae or other plants.
Opportunities for carbon sequestration
Soil sequestrationChange management practices on cropped and grazed land to sequester carbon in soils.
Restoration of mulga countryRestore degraded, low value grazing country in arid Australia.
PlantationsEstablish plantations for timber production and specifically to sequester carbon.
Pre-1990 eucalypt forestsTimber harvesting and other human disturbances are reduced to allow forests to sequester the maximum amount of carbon.
Source: Garnaut, 2008, The Garnaut Climate Change Review.

Passenger transport

Fuel use in road transport is a significant source of greenhouse gas emissions and passenger cars make up a large component of this. In terms of kilometres travelled for every unit of energy used, buses are the most energy efficient mode of passenger transport, followed by heavy rail and motor cycles. Passenger cars and inland ferries are the least efficient (Endnote 53).

Carbon dioxide emissions in new motor vehicles are falling. According to the National Transport Commission, in 2008, the national average carbon emissions from new passenger and light commercial vehicles was 222 g/km, a 12% reduction from 2002 (Endnote 54).

In Australia in 2009, the most common factor considered when buying a motor vehicle was the purchase price (53%). While fuel economy/running costs was the second most common factor (41%), environmental impact/exhaust emissions were considered by only 4% (Endnote 55). This is reflected in the fact that in 2008 only 1% of car sales in Australia were “green” cars (Endnote 56), compared to 11% in the UK (Endnote 57).

TRAVEL TO WORK: SELECTED MODES, 2009

Main form of transport used on usual trip to work or full-time study, March 2009%

Private motor vehicle79.6
Public transport14.0
Bicycle1.5
Walk4.0
Other0.9

Source: ABS, 2009, Environmental Issues: Waste Management and Transport Use, Mar 2009 (ABS cat. no. 4602.0.55.002).

In 2009, over three quarters (80%) of Australians used private motor vehicles as the main method of travel to work or study, compared to only 14% who used public transport (Endnote 58). However, the proportion of trips on public transport has risen since 1996, particularly in Victoria.
USE OF PUBLIC TRANSPORT, 1996 AND 2009
Graph: Use of public transport, 1996 and 2009
Note: Public transport used as the main form of transport on usual trip to work or full-time study
Source: ABS, 2006, Environmental Issues: People’s Views and Practices March 2006 (cat. no. 4602.0);
ABS, 2009, Environmental Issues: Waste Management and Transport Use, March 2009 (ABS cat. no. 4602.0.55.002).


Lack of public transport services at the right time and complete lack of services continue to be the main reasons why people in Australia do not use public transport. In 2009, over one half (52%) of people not taking public transport cited either of these reasons. The convenience/comfort/privacy of using a motor vehicle and travel time were the next most common reasons for not using public transport (22% and 18%, respectively) (Endnote 59).

Relatively few people usually walked or cycled to their place of work or study (4% and 1%, respectively). The proximity of home (64%) and exercise and health (50%) were the two most commonly reported reasons why people walked or cycled. Only 7% cited environmental concerns as a reason for walking (Endnote 60).

REASONS FOR NOT TAKING PUBLIC TRANSPORT TO WORK OR STUDY, MARCH 2009
Graph: Reasons for not taking public transport to work or study, March 2009
Note: Major reasons given by respondents
Source: ABS, 2009, Environmental Issues: Waste Management and Transport Use March 2009 (cat. no. 4602.0.55.002).


The renewable energy industry

In 2007–08, Australia produced 290 petajoules (PJ) of renewable energy, equivalent to 5% of Australia’s total primary energy consumption of 5,572 PJ. Bagasse (sugar cane waste) was the largest source of renewable energy (39%), followed by wood and wood waste (33%) and hydroelectricity (15%). Other biofuels, wind and solar (including solar hot water) contributed 13%.

AUSTRALIAN PRODUCTION OF RENEWABLE ENERGY (a)

2001-02
2007-08
%
(PJ)
(PJ)
Increase

Bagasse
91.7
111.9
22
Other biofuels (b)
10.1
17.6
74
Hydroelectricity
57.5
43.4
-25
Solar hot water
2.7
6.5
141
Wind and solar photovoltaic
0.6
14.6
2333
Wood and woodwaste
95.0
96.0
1
Total
257.6
290.0
13

(a) Electricity and heat.
(b) Includes biogas, black liquor, crop and municipal waste.
Source: Australian Bureau of Agricultural and Resource Economics (ABARE), 2009, Energy statistics – historical, Table A, <http://www.abare.gov.au/publications_html/data/data/data.html>, last viewed November 2009;
ABARE, 2009, Energy in Australia 2009.


Over the six year period from 2001–02 to 2007–08, production of renewable energy increased by 13% (Endnote 61). Very rapid growth rates in wind and solar electricity were from low bases and have been largely offset by falling hydroelectricity generation.

According to the Australian Bureau of Agricultural and Resource Economics, in the six months to April 2009, ten electricity generation projects were completed in Australia, four of which were renewables (two wind and two biomass). The new renewable energy projects have a generating capacity of 310 megawatts (MW), or 16.5% of additional generating capacity commissioned (Endnote 62).

Wind farms figure very prominently in renewable electricity projects. The recently completed Waubbra Wind Farm in Victoria is the largest in the southern hemisphere, consisting of 128 wind turbines with a generation capacity of 192 MW (Endnote 63).

In April 2009, renewable electricity projects represented 16% of capacity of all electricity generation projects (Endnote 64) at an advanced stage of development (Endnote 65). In addition, the number of renewable electricity projects at less advanced stages of development is considerable and the combined capacity of these projects constitutes 41% (9,408 MW) of all less advanced electricity generation projects (Endnote 66).

Sources of energy in homes

Energy consumption in the residential sector is a significant contributor to greenhouse gas emissions in Australia due to the heavy reliance on fossil fuels, notably coal, to produce electricity.

Hot water systems and space heating account for the majority of energy used in most households. In March 2008, electricity was the main energy source for hot water systems (46%), space heating (35%), ovens (75%) and cooktops (56%) (Endnote 67).

Compared to electricity generated from coal, natural gas produces substantially less carbon dioxide emissions (Endnote 68). Gas is the second most common source of energy for Australian households and was used in more than half of households (61%) in March 2008, particularly in the gas producing states of Victoria and Western Australia (Endnote 69).

Use of renewable energy is still uncommon in Australian homes. Solar energy use has increased from 5% in 2002 to 8% in 2008. It is used primarily for heating water (Endnote 70).
MAIN SOURCES OF ENERGY IN DWELLINGS, 2008
Graph: Main sources of energy in dwellings, 2008
Note: Only includes the five most common sources of energy used in residential dwellings.
Source: ABS, 2008, Environmental Issues: Energy Use and Conservation March 2008 (cat. no. 4602.0.55.001).


The use of solar hot water varies considerably between regions. In the Northern Territory and Western Australia, 54% and 21% of households, respectively, had solar hot water, compared to the national figure of 7% (Endnote 71). These higher proportions reflect numerous factors but especially high levels of solar radiation in these states and the larger proportion of remote communities lacking access to cheap mains electricity.

SOLAR HOT WATER HEATING, USE IN DWELLINGS
Graph: Solar hot water heating, use in dwellings
Source: ABS, 2008, Environmental Issues: Energy Use and Conservation March 2008 (cat. no. 4602.0.55.001).


GreenPower is a government renewable energy accreditation program. GreenPower schemes enable consumers to pay a premium for electricity generated from sources like mini hydro, wind power and biomass which produce no net greenhouse gas emissions.

The schemes have been operating for over ten years in New South Wales, Victoria, Queensland, Western Australia, South Australia and the Australian Capital Territory. In the March quarter 2009, there were approximately 984,000 GreenPower customers in Australia, a substantial increase from 138,879 customers in March 2005 (Endnote 72).

More than half of all households (52%) were aware of GreenPower in 2008 (including 5% already paying for GreenPower). This was a large increase compared with 2005 and 1999 when 29% and 19% respectively were aware of GreenPower (Endnote 73).

In 2008, one-third of households that were aware of GreenPower stated they were willing to pay more to support the scheme, up from 23% in 2005 (Endnote 74).

AWARENESS OF GREENPOWER SCHEME, MARCH 2008

NSW
Vic.
Qld
SA
WA
ACT
Aust.
%
%
%
%
%
%
%

Already paying for GreenPower
5.0
7.1
5.3
5.8
*1.0
4.9
5.3
Aware of GreenPower scheme
48.9
52.9
38.4
45.3
38.4
65.9
46.7
Not aware of GreenPower scheme
42.2
35.6
53.1
45.4
59.0
26.3
44.4
Did not know
3.9
4.4
3.2
3.5
1.6
*2.9
3.6
Total
100
100
100
100
100
100
100

* estimate has a relative standard error of 25% to 50% and should be used with caution
Note: Data covers only states and territories that are participating in the National Green Power Accreditation Program.
Source: ABS, 2008, Environmental Issues: Energy Use and Conservation March 2008 (cat. no. 4602.0.55.001).

WILLINGNESS TO PAY EXTRA PER ANNUM ON GREENPOWER ELECTRICITY, MARCH 2008

NSW
Vic.
Qld
SA
WA
ACT
Aust.
%
%
%
%
%
%
%

Willing to pay extra
30.9
30.9
34.6
30.9
37.6
36.1
32.5
Not willing to pay extra
57.7
56.0
52.3
55.3
53.1
54.0
55.4
Did not know
11.4
13.1
13.1
13.8
9.3
9.9
12.1
Total
100
100
100
100
100
100
100

Note: Data covers only states and territories that are participating in the National Green Power Accreditation Program.
Source: ABS, 2008, Environmental Issues: Energy Use and Conservation March 2008 (cat. no. 4602.0.55.001).


Energy efficiency in homes

Residential buildings are responsible for a significant proportion of Australia’s emissions, in both construction and use.

In 2007–08, most Australians (88%) reported that they take steps to limit their electricity use. The main reasons people gave for not taking steps to limit electricity use was that their electricity consumption was already low enough (33%) and that they had not thought about saving electricity (27%) (Endnote 75).

Electrical appliances account for around 30% of energy use in the home (Endnote 76). In 2008, more than three-quarters (77%) of all households had a heater, over two-thirds (67%) had a cooler (i.e. air conditioner or evaporative cooler) and more than half (56%) had a clothes dryer. Nearly half of households had dishwashers (45%) and more than one-third had separate freezers (37%) (Endnote 77).

When buying new appliances, energy efficiency was the most commonly reported factor which influenced the decision to buy a refrigerator (72%) and air conditioner (74%). Purchase price was the second most commonly reported factor for these appliances (68% and 63% respectively) (Endnote 78).
HOUSES WITH INSULATION
Graph: Houses with insulation
Source: ABS, 2008, Environmental Issues: Energy Use and Conservation March 2008 (cat. no. 4602.0.55.001).


In 2008, 61% of households reported having insulation in their dwelling, up from 52% in 1994. The main reason given for having insulation was to improve comfort (83% of households installing insulation). While only a small proportion (4%) of households reported that they had installed insulation primarily to save energy, the installation of insulation for whatever reason leads to lower energy use (Endnote 79).

Putting a price on carbon

Sir Nicholas Stern has described climate change as the greatest example of market failure we have ever seen (Endnote 80). The failure to put a price on emissions has led to excessive emissions and the risk of dangerous climate change (Endnote 81).

Putting a price on carbon (Endnote 82) through the introduction of emissions trading has been proposed as one way to reduce emissions. Under a cap and trade scheme, permits are required to emit greenhouse gases into the atmosphere. These permits can be bought and sold but the government is able to place a cap on total emissions by limiting the number of permits issued (Endnote 83).

Industries in Australia most likely to be affected by putting a price on carbon are those with: 1) a high emissions intensity; 2) limited or negligible access to substitutes to reduce their emissions intensity; 3) exposure to increased costs under an emissions trading scheme; 4) limited capacity to pass on the emissions price.

In July 2008, the Australian Government released a Green Paper on a Carbon Pollution Reduction Scheme which included options and preferred approaches relating to imposing a limit on how much carbon pollution industry can emit.

Although in the proposed scheme the cost of purchasing emissions permits would rest with certain emissions-intensive industries, such as electricity generators, the cost is expected to be passed down the supply chain.

Consumers would, therefore, pay more for a range of goods and services, particularly emissions-intensive goods and services. Low income households are particularly vulnerable to price increases as they spend a greater proportion of their incomes on items that are more likely to be impacted by higher energy prices, such as food, petrol, electricity and gas (Endnote 84).

EXPENDITURE ON SELECTED ITEMS AS A PROPORTION OF DISPOSABLE INCOME

Graph: Expenditure on selected items as a proportion of disposable income
Source: ABS, data available on request (Household Expenditure Survey, 2003–04).


Some households already struggle to purchase necessities. In 2003–04, almost 9% of households with low income and low net worth reported they were unable to heat their home. Twelve percent went without meals and 38% could not pay utility bills on time (Endnote 85).

SELECTED INDICATORS OF FINANCIAL STRESS, 2003-04

Low economic resources households (a) (%)
Other households (%)

Unable to heat home
8.9
1.2
Went without meals
11.8
1.8
Could not pay bills on time
37.8
11.5
Total households ('000)
1 050.6
6 685.2

(a) Households simultaneously in both the lowest three income deciles and the lowest three net worth deciles.
Source: ABS, data available on request (Household Expenditure Survey, 2003–04).

People living in rural or outer suburban areas may also be disproportionately affected. Higher transport costs in these areas will also be reflected in higher product prices, including food. Those who need to drive long distances to access services will be particularly vulnerable to rising fuel prices (Endnote 86).

The effects of higher fuel and power prices can be offset by a range of measures. For example, under an emissions trading scheme, revenues gained from the sale of emission permits to industry can be used to compensate households, improve access to public transport and assist households to use more energy efficient products and motor vehicles.
IMPACTS AND ADAPTATION

“Mitigation will come too late to avoid substantial damage from climate change”.
Garnaut, 2008 (Endnote 87).

Within the next few decades, it is likely that Australian households, communities and businesses will have to take actions to adapt to the impacts of climate change that cannot be avoided by mitigation (Endnote 88).

Impacts affecting society, environment and the economy

The impacts of climate change will affect the environment, society and the economy. The vulnerability of these systems will vary between regions and sectors depending on exposure to changes in the climate, sensitivity to those changes and capacity to adapt.

This section explores some of the areas considered most vulnerable to the impacts of climate change:
  • Water
  • Agriculture
  • Biodiversity
  • Coastal settlements
  • Human health

While some households and businesses are already taking actions to adapt to a changing climate, some are limited in their capacity to adapt and, therefore, may be more vulnerable.

Water

In 2004–05, the distribution of water consumption in the Australian economy was:
  • 65% by agriculture;
  • 11% by households;
  • 11% water supply industry (including losses in delivery system);
  • 3% by manufacturing;
  • 10% by other industries (e.g. mining, service industries) (Endnote 89).

Lower rainfall and increases in evaporation will reduce runoff and stream flow in many parts of Australia, including many major cities and irrigation areas. For example, in one study a temperature increase of 1 to 2°C is projected to result in a 7 to 35% decrease in Melbourne’s water supply and a 0 to 25% decrease in flow in the Murray-Darling Basin (Endnote 90).

Projections suggest that across Australia the number of drought months will increase by up to 20% by 2030. By 2070, drought months are projected to increase by up to 40% in eastern Australia and by up to 80% in south-western Australia (Endnote 91).

Dams

Dams have been built in Australia since the late-1800s to provide a reliable water resource for irrigated agriculture, urban water needs and hydro-electric power generation (Endnote 92).

At the start of the 20th century the combined storage capacity of all large dams was 250 GL. This grew to 9,540 GL by 1950 and 83,853 GL in 2005 (Endnote 93). Australia's 500 large dams have a total capacity equivalent to 4,000 kilolitres (kL) per person.

TOTAL STORAGE CAPACITY OF LARGE DAMS
Graph: Total storage capacity of large dams
Note: A large dam is defined as having a height of greater than 15 metres (m), or as greater than 10 m but meeting other size criteria.
Source: ABS, 2006, Water Account Australia 2004–05 (cat. no. 4610.0).


In addition, there are many thousands of farm dams throughout Australia. Australia’s high per capita storage capacity is needed to sustain agricultural production and potable water supplies for human use during long dry periods.

LARGE DAM STORAGE LEVELS

Graph: Large dam storage levels
Source: ABS, 2006, Water Account Australia 2004–05 (cat. no. 4610.0).


Drought conditions were reflected in an 18% fall in the water stored in large dams between 2002 and 2005. On 1 July 2002, storage levels were at 48,683 GL, falling to 39,959 GL by 30 June 2005 (Endnote 94).

Comprehensive current data for all of Australia is not available. However, total public storage in the Murray-Darling Basin at the end of October 2009 was only 6,450 GL, or 28% of capacity (Endnote 95).

Water management on farms

Australia’s agriculture industry is particularly dependent on irrigation water to sustain production. Whilst most agricultural water is used for irrigation of crops and pasture, water is also used for livestock drinking and washing down dairy sheds.

In 2004–05, a third of all farms carried out water-related management activities, spending a total of $314 million in that year (Endnote 96).

The most commonly reported water management activities were: earthworks, drains and water pumping; tree and shrub maintenance; and removing stock from waterways (Endnote 97).

Water issues included surface and groundwater availability, excess nutrients, clarity, toxicity and others. Of these, water availability was the water issue most frequently reported by farmers (Endnote 98).
FARMS IDENTIFYING WATER ISSUES AND WATER ACTIVITIES, 2004-05
Graph: Farms identifying water issues and water activities, 2004-05
Source: ABS, 2007, Natural Resource Management on Australian Farms 2004–05 (cat. no. 4620.0).


EXPENDITURE BY THE AGRICULTURE INDUSTRY FOR WATER MANAGEMENT, 2004-05

NSW
Vic.
Qld
SA
WA
Tas.
NT
ACT
Aust.

Total expenditure
128
^51
^85
^18
25
^5
*1
^-
314
($ million)
Average expenditure per farm ($)
9 501
^5 151
^9 241
4 836
5 095
3 833
*18 474
^3 405
7 351

^ Estimate has a relative standard error of 10% to less than 25% and should be used with caution.
* Estimate has a relative standard error of 25% to less than 50% and should be used with caution.
Source: ABS, 2007, Natural Resource Management on Australian Farms 2004–05 (cat. no. 4620.0).


Inland waterways and wetlands

Climate change and, in particular, reduced rainfall, increased drought, more intense rainfall events, sea level rise and warming of the water column will impact on inland waterways and wetlands in many ways, including:
  • reduced river flows and changes in seasonality of flows;
  • changes in species composition and community structure (such as loss of cool adapted aquatic species);
  • reduced area available for waterbird breeding;
  • sea level rise resulting in saltwater intrusion into freshwater bodies; and
  • changes in water quality, eutrophication (Endnote 99) levels and incidence of blue-green algae outbreaks (Endnote 100).

In the interest of maintaining the health of rivers, a number of states and territories are allocating and providing water to the environment – generally known as ‘environmental flows’.

Without sufficient flows water-dependent ecosystems may lose their capacity to provide for environmental and other public benefits outcomes. Such losses can be difficult or costly to overturn and, in some cases, may be irreversible (Endnote 101).

The ABS Water Account Australia 2004–05 presented information on water released for the purpose of the environment in accordance with specific environmental regulations. This has been termed environmental provisions; it does not represent all environmental flows, but only the volume of water released by water suppliers. Other methods of providing water to the environment include placing limits and rules on licences for water extraction and strategic management of flows and water quality.

In 2004–05, 1,005 GL of water was supplied to the environment through environmental provisions. This is an increase of 119% across Australia since 2000–01. States with large increases were Queensland, Victoria and Tasmania (Endnote 102).

More recently, the Water Act 2007 (Cwlth) established a management system for the Australian government’s $3.1 billion Restoring the Balance in the Murray-Darling Basin program and $5.8 billion Sustainable Rural Water Use and Infrastructure program (Endnote 103).

ENVIRONMENTAL PROVISIONS

NSW/ACT
Vic.
Qld
SA
WA
Tas.
NT
Aust.
GL
GL
GL
GL
GL
GL
GL
GL

2000-01
200.5
253.2
4.5
0.9
-
0.4
-
459.4
2004-05
127.2
373.9
383.6
0.7
-
118.7
1.1
1 005.3

– Nil or rounded to zero (including null cells).
Source: ABS, 2006, Water Account Australia 2004–05 (cat. no. 4610.0).


Household water use and conservation

After years of low rainfall, Australian households are adapting to reduced water availability. Over the period 2000–01 to 2004–05, household water use per person fell 14%, from 120 kL to 103 kL. Tasmania was the only state to record an increase (Endnote 104).

Decreased household use reflects water restrictions in most states and territories since 2002, government incentives to reduce water use and improve water use efficiency, and voluntary conservation of water by households.

HOUSEHOLD WATER CONSUMPTION PER PERSON

Graph: Household water consumption per person
(a) Includes unlicensed water use from garden bores.
Source: ABS, 2006, Water Account Australia 2004–05 (cat. no. 4610.0).

From 1994 to 2007, the proportion of households with water conservation devices more than doubled (Endnote 105).

HOUSEHOLDS WITH WATER CONSERVATION DEVICES

Graph: Households with water conservation devices
Source: ABS, 2007, Environmental Issues: People’s Views and Practices March 2007 (cat. no.4602.0).


Whilst mains/town water is overwhelmingly the principal source of water for Australian households, (93% connected in March 2007), households are reducing their reliance on mains water by increasing their use of grey water and rainwater tanks (Endnote 106).

RAINWATER TANKS AS A SOURCE OF WATER FOR HOUSEHOLDS

Graph: Rainwater tanks as a source of water for households
Source: ABS, 2007, Environmental Issues: People’s Views and Practices March 2007 (cat. no.4602.0).


In 2007, nearly one-fifth (19%) of all households sourced water from a rainwater tank, up from 16% in 2001 (Endnote 107).

Agriculture

Agriculture is an important part of the Australian economy. In 2007–08, the gross value of agricultural production was $43.3 billion (Endnote 108), and in 2008–09, 318,000 people were employed in the agriculture industry (Endnote 109).

Australia's agricultural businesses are engaged mainly in beef cattle farming, dairy cattle farming, sheep farming, grain growing, or a mixture of two or more of these activities. Farm exports account for around 15% of total merchandise exports (Endnote 110), with products such as beef, wheat, and skim milk powder contributing significantly to global markets.

GROSS VALUE OF AGRICULTURAL COMMODITIES PRODUCED, 2007-08


Climate change is likely to affect agriculture in a number of ways:
  • changes in rainfall and temperature will affect crop production;
  • changes in the quantity and quality of pasture as well as temperature increases will affect the productivity of the livestock industries;
  • severe weather events (e.g. bushfires and flooding) will affect crop yields and stock;
  • changes in temperature are expected to alter the incidence and occurrence of pests and disease; and
  • where there is adequate moisture, increased concentrations of CO2 will increase growth in some plants (Endnote 111).

Many Australian farmers are already taking actions to adapt to what they perceive as a changing climate. In 2006–07, 66% of Australian agricultural businesses reported that the climate affecting their holding had changed and of this group 75% reported that they had changed management practices as a result of this perceived change (Endnote 112).

The most commonly reported perceived change in climate affecting the holding was a change in rainfall patterns (92%) followed by more extreme weather events (74%) and warmer temperatures (50%) (Endnote 113).

The most commonly reported impact on the holding was a decreased level of production (89%) followed by an increased frequency or extent of pests, weeds or disease (56%).

In contrast, a small proportion of agricultural businesses reported a decreased frequency or extent of pests, weeds or disease (20%) and an increased level of production (15%) (Endnote 114).
AGRICULTURAL WATER USE ON AUSTRALIAN FARMS

Graph: Agricultural water use on Australian farms
Note: Northern Territory data too small to display.
Source: ABS, 2008; 2008; 2009, Water Use on Australian Farms 2005–06; 2006–07; 2007–08 (cat. no. 4618.0)


From 2005–06 to 2007–08, agricultural water use on Australian farms fell 40% (from 11,689 GL to 6,989 GL) due to the continuing unavailability of water as a result of the drought. The largest declines occurred in NSW (61%), Victoria (44%) and Queensland (21%) (Endnote 115).

GROSS VALUE OF IRRIGATED AGRICULTURAL PRODUCTION, SELECTED PRODUCTS

Graph: Gross value of irrigated agricultural production, selected products
Note: Year refers to financial year eg. 2001 refers to 2000-01.
Source: ABS, 2009, Experimental Estimates of the Gross Value of Irrigated Agricultural Production, 2000–01 to 2006–07 (cat. no. 4610.0.55.008).


In 2006–07, irrigated agricultural land comprised less than 0.5% of all agricultural land in Australia, yet the gross value of irrigated agricultural production (GVIAP) was $12,319 million. This represented 34% of the total gross value of agricultural production, up from 28% in 2000–01 (Endnote 116).

The GVIAP generated for each megalitre of water applied (GVIAP/ML) can vary significantly between different agricultural activities and over time. The product groups with the highest GVIAP/ML in 2006–07 were: nurseries, cut flowers and cultivated turf, ($16,470), vegetables ($6,104), and fruit and nuts ($4,493). The product with the lowest GVIAP/ML was rice ($230) (Endnote 117). Reductions in water availability could see a decline in low GVIAP/ML activities to higher ones.

GVIAP PER MEGALITRE OF WATER APPLIED, SELECTED PRODUCTS
Graph: GVIAP per megalitre of water applied, selected products
(a) Nurseries, cut flowers and cultivated turf.
(b) Vegetables for human consumption or seed.
Source: ABS, 2009, Experimental Estimates of the Gross Value of Irrigated Agricultural Production, 2000–01 to 2006–07 (cat no. 4610.0.55.008).


Murray-Darling Basin

The Murray-Darling Basin (MDB) covers around 14% of Australia’s land area (Endnote 118) and is of special importance to Australia’s agricultural production. In 2005–06, the gross value of agricultural production (GVAP) in the Basin was worth $15 billion or 39% of Australia’s total GVAP ($38.5 billion) (Endnote 119).

Agriculture dominates land use in the Basin. In 2006, 10% of the people employed in the Basin worked in agriculture, compared to only 3% Australia wide (Endnote 120).

Industries (including agriculture) and households in the MDB accounted for just over half (52%) of Australia’s total water consumption in 2004–05. In terms of agricultural water consumption, the MDB is even more dominant. In 2005–06 the MDB used two-thirds (66%) of Australia’s total agricultural water consumption (Endnote 121).

Irrigated agriculture in the MDB generated $4.6 billion or 44% of Australia’s gross value of irrigated agricultural production in 2005–06. Dairy farming generated $938 million, fruit and nuts $898 million, cotton $797 million and grapes $722 million (Endnote 122).

Irrigated agriculture in the MDB is one area for which the impacts of climate change are anticipated to be large. Lower average rainfall and higher average temperatures are expected to significantly reduce water availability in the Basin, impacting on crop yields and the quantity of water available for irrigation.

Without effective mitigation, by 2100, the economic production of irrigated agriculture in the Basin is projected to fall by 92% (Endnote 123).
AGRICULTURE WATER CONSUMPTION, 2005-06

Murray-Darling Basin
Australia
MDB as a proportion of Aust.
GL
GL
%

Dairy farming (a)
1 287
1 893
68
Pasture for other livestock (b)
1 284
2 042
63
Rice
1 252
1 253
100
Cereals (excl. rice)
782
894
88
Cotton
1 574
1 735
91
Grapes
515
633
81
Fruit (excl. grapes)
413
630
66
Vegetables
152
431
35
Other agriculture (c)
461
2 178
21
Total Agriculture
7 720
11 689
66

(a) Includes irrigated pasture for grazing, hay and seed; livestock drinking; and shed washdown.
(b) Includes irrigated pasture for grazing, hay and seed.
(c) Includes other broadacre crops, nurseries, livestock (other than dairy) drinking.
Note: Components may not add to total due to rounding.
Source: ABS, 2008, Water and the Murray-Darling Basin, A Statistical Profile 2000–01 to 2005–06 (cat. no. 4610.0.55.007).


One way farmers can adapt to reduced water availability is by improving their water use efficiency. It is estimated that currently between 10 and 30 per cent of the water diverted from the rivers into irrigation systems is lost before it reaches the farm, and up to 20 per cent of the delivered water may be lost in on-farm distribution channels (Endnote 124).

Approximately two-thirds of irrigators in the MDB changed their water management practices during 2004–05. In 2004–05, the most common changes to irrigation practices in the MDB (as a proportion of total MDB irrigated farms) were:
  • adopting more efficient irrigation techniques (35%)
  • undertaking more efficient irrigation scheduling (27%)
  • reducing area under irrigation (20%)
  • laser levelling (17%)
  • purchasing extra irrigation water (16%) (Endnote 125).

Less common practices included: introducing reuse or recycled irrigation water (11%), installing soil moisture sensors (9%), selling irrigation water (8%), increasing the area under irrigation (8%), and installing piping or covering open channels to reduce water loss (7%) (Endnote 126).

Biodiversity

Australia is one of the world’s most biologically diverse countries. It is home to more than one million species, many of which are unique.

Biodiversity is particularly vulnerable to climate change because it has relatively low adaptive capacity compared to other sectors (Endnote 127).

In addition to climate change, biodiversity in Australia is under pressure from a range of other factors such as land clearing, pollution, weeds and pests. It is, therefore, difficult to know with certainty exactly what impact climate change has had on biodiversity to date. Nonetheless, numerous changes in biodiversity have been observed that are consistent with climate change. For example:
  • changes in geographic ranges including expansion of animal and plant species into higher elevations and southerly latitudes;
  • earlier arrival and later departure of migratory bird species;
  • declines in populations of mountain pygmy possums associated with declining snow cover;
  • expansion of rainforest in the NT, QLD and NSW;
  • altered fire regimes; and
  • more intense and frequent coral bleaching events (Endnote 128).

The capacity of natural systems to adapt to climate change will be improved if other pressures, such as land clearing, lack of environmental flows and pollution, can be eased. The resilience of natural systems can also be improved by expanding reserve systems and promoting conservation on private land (Endnote 129).


THE GREAT BARRIER REEF
The Great Barrier Reef is among the largest and most spectacular coral reef ecosystems in the world. A World Heritage Area, it is home to many organisms including six species of marine turtles, 24 species of seabirds, more than 30 species of marine mammals, 350 coral species, 4000 species of molluscs and 1500 fish species. Coral reefs are highly vulnerable to human-induced climate change.

Over the last 30 years, the waters of the Great Barrier Reef have increased in temperature by 0.4°C. This has made the Reef more susceptible to heat stress during short periods of warm sea temperature. As a result, episodes of mass coral bleaching have increased in frequency and intensity.

Over the last 25 years, heat stress has resulted in six episodes of coral bleaching within the Reef. In 1998, 50% of the coral reefs within the Great Barrier Reef Marine Park were affected by coral bleaching and in 2002, another mass coral bleaching event affected 60% of the coral reefs. About 5 to 10% of the corals affected during these events died.

Ocean chemistry has also been affected by climate change. Ocean pH has decreased by 0.1, that is, the ocean is becoming more acidic. As CO2 concentrations increase in the atmosphere, increased amounts of CO2 enter the ocean where it combines with water to produce a weak acid which limits the rate of coral growth. While the impacts of ocean acidification are not yet fully understood, there is consensus that ocean acidification poses a serious threat to coral reefs.

Increasing water temperatures and ocean acidity are putting this unique ecosystem at risk. Even with effective mitigation, it is expected that, by 2100, mass coral bleaching will be twice as common as it is today. Without mitigation, by 2100, the Great Barrier Reef will be destroyed.

Damage to or destruction of the Great Barrier Reef will have serious implications for the Queensland economy. Tourism is an important part of Queensland’s economy and a substantial proportion of tourism in Queensland is related to the existence of the Great Barrier Reef. It is estimated that the reef interested tourism economy contributes more than $2 billion each year to Queensland’s Gross State Product. Tourism in the Tropical North region is particularly dependant on the reef with over 90% of interstate and international visitor nights associated with interest in the reef.

Source: Hoegh-Guldberg, O, and Hoegh-Guldberg, H, 2008, Garnaut Climate Change Review: The impact of climate change and ocean acidification on the Great Barrier Reef and its tourist industry.
Eco-tourism

Australia’s natural landscapes underpin much of Australia’s international and domestic tourism.

Each year, millions of domestic and international visitors in Australia participate in nature activities such as:
  • visiting national parks, wildlife parks, zoos, aquariums, botanical gardens and public gardens;
  • bushwalking;
  • whale and dolphin watching; and
  • snorkelling and scuba diving (Endnote 130).

Total expenditure by domestic visitors who participated in nature activities was approximately $12 billion in 2008 (Endnote 131).

Two-thirds (65%) of the 3.36 million international visitors to Australia in 2008 participated in nature activities. These visitors spent $20.2 billion (Endnote 132).

NUMBER OF VISITORS WHO PARTICIPATED IN NATURE ACTIVITIES, 2004 TO 2008

2004
2006
2008

Domestic overnight
No. of visitors (million)
12.62
13.15
12.94
Share of total (%)
17
18
18
Domestic day
No. of visitors (million)
11.01
12.44
12.37
Share of total (%)
8
9
9
International
No. of visitors (million)
3.02
3.43
3.36
Share of total (%)
63
67
65

Source: Tourism Research Australia, 2009, Nature Tourism in Australia 2008.


Some natural attractions would be significantly affected by unmitigated climate change, particularly the Great Barrier Reef. More generally, beaches are in danger of increasing storm damage and inundation. Ski fields will suffer from reductions in snow cover, average season lengths and peak snow depths (Endnote 133).

Popular tourist destinations may become less appealing if they face climate change related impacts such as increased incidence of bushfires, increased ultraviolet radiation, increased exposure to disease and increased extreme weather events (e.g. flooding, storm surges, heatwaves, cyclones and droughts). Climate change is also expected to lead to increased costs associated with increased need for repair, maintenance and replacement of tourist infrastructure as well as development of alternative attractions (Endnote 134).

Coastal settlements

“Australia’s coastal zone is a significant national environmental asset that is also fundamentally important to our lifestyle and economy”.
House of Representatives Standing Committee on Climate Change, Water, Environment and the Arts (Endnote 135).

Coastal communities, their infrastructure and resources are vulnerable to a number of climate change impacts.
Sea level rise is likely to result in:
  • increased risk of inundation during storm surges;
  • increased coastal erosion and recession;
  • loss of wetlands and mangroves saltwater intrusion into freshwater sources; and
  • loss of wetlands (Endnote 136).

Extreme weather events will also impact upon coastal areas. For example tropical cyclones are expected to become more intense in northern Australia (Endnote 137).

The majority of Australians (over 80%) live within the coastal zone. About 711,000 addresses are within three kilometres of the coast and less than six metres above sea level (Endnote 138), and coastal settlements are continuing to grow. In 2007–08, outside capital cities, the largest population growth generally occurred along the Australian coast. Several local government areas on the Queensland coast had large population increases, such as the Gold Coast (up 13,000 people), Sunshine Coast (9,000), Cairns (6,000) and Townsville (5,000) (Endnote 139). Other coastal centres experiencing rapid growth included Seaside Tweed in NSW, Mandurah and Busselton in Western Australia and Victor Harbour in South Australia. Areas of population decline occurred mainly in inland rural areas.

POPULATION DENSITY, AUSTRALIA, JUNE 2008
Map: Population density, Australia, June 2008

Source: ABS, 2009, Regional Population Growth Australia 2007–08 (cat. no. 3218.0).
Human health

“Climate change is a significant and emerging threat to public health, and changes the way we must look at protecting vulnerable populations”.
World Health Organisation (Endnote 140).

In Australia, some health impacts attributable to climate change will be direct, such as death and disease associated with heatwaves and natural disasters. Others will occur indirectly, such as increases in mosquito borne diseases due to changes in mosquito population range and activity (Endnote 141).

Climate change will also impact upon food, water and air quality, which are the most fundamental determinants of health (Endnote 142).

Australians most ‘at risk’

Some communities in Australia are more vulnerable to climate change than others. This reflects differences in exposure to risks as well as adaptive capacity. People living in remote areas, people on lower incomes, those with poor housing, the elderly and the sick are among the most vulnerable (Endnote 143).

Torres Strait islanders and remote indigenous communities are particularly vulnerable because of their relative isolation and limited access to support facilities (Endnote 144).

Climate change is expected to result in substantial increases in extreme hot weather. If no attempt is made to mitigate climate change, by 2100 the number of days over 35°C each year is projected to rise from 9 to 27 in Melbourne, 1 to 21 in Brisbane and most dramatically from 9 to 312 in Darwin (Endnote 145).

Extreme hot weather has serious impacts on health, including heat-related deaths from heatwaves. The IPCC projects increased frequency of heatwaves over this century as “very likely”, with an associated increased risk of heat related deaths (Endnote 146).


CLIMATE CHANGE AND HEALTH
In 2008, the Garnaut Review identified the main health risks climate changes poses in Australia. The risks are many and varied and include:

· impacts of severe weather events (floods, storms, cyclones);

· impacts of temperature extremes, including heatwaves;

· mosquito-borne infectious diseases (e.g. dengue and Ross River virus);

· food-borne infectious diseases (e.g. those due to Salmonella and Campylobacter);

· water-borne infectious diseases and health risks from poor water quality;

· diminished food production and higher prices, with nutritional consequences;

· increases in air pollution (e.g. from bushfires);

· changes in production of aeroallergens (spores, pollens), with the potential to exacerbate asthma and other diseases; and

· mental health consequences and the emotional cost of social, economic and demographic dislocation (e.g. in parts of rural Australia, and through disruptions to traditional ways of living in remote Indigenous communities).

Source: Garnaut, 2008, The Garnaut Climate Change Review;
Bambrick et al., 2008, The impacts of climate change on three health outcomes: temperature-related mortality and hospitilisations, salmonellosis and other bacterial gastroenteritis, and population at risk from dengue.


The elderly are particularly vulnerable to extreme hot weather. A combination of factors such as chronic illness, disability, prescribed medication and social isolation reduce the capacity of elderly individuals to cope during heat waves (Endnote 147).

In this regard, the growing proportion of elderly people in Australia is of concern. Between 30 June 1989 and 30 June 2009, the proportion of Australia's population aged 65 years and over increased from 11% to 13%. During the same period, the proportion of the population aged 85 years and over doubled, from 0.9% to 1.8% (Endnote 148).

This trend is expected to continue. Projections by the ABS, based on certain assumptions about future levels of fertility, mortality and net overseas migration, indicate that by 2056 the proportion of Australians aged 65 and over may reach around 23% to 25%, while the proportion aged over 85 years may reach 5% to 7% (Endnote 149). This trend is evident in the population pyramid shown.

POPULATION STRUCTURE, AGE AND SEX – AUSTRALIA – 2006 AND 2056
Image: Population structure, age and sex - Australia - 2006 and 2056
Source: ABS, 2009, data available on request.


SUMMARY

In 2007–08, almost three out of four Australians were concerned about climate change (Endnote 150).

Climate change and strategies to mitigate and adapt to it will affect many facets of Australian society, including the sources of energy used to power homes and industry, town planning, modes of transport and how farmers manage the land.

If not already, over the next few decades many Australians will experience impacts associated with climate change. Reduced rainfall in many areas will affect urban water supplies and agriculture.

Biodiversity, already under pressure, is particularly vulnerable to the impacts of climate change and unique ecosystems such as the Great Barrier Reef are at risk. Coastal erosion, storm surges and increased risk of inundation will affect coastal communities. Human health will also be affected.

Improving our understanding of the complex relationships between climate and natural and human systems will be necessary to better identify and manage the many impacts of climate change on businesses, industry, individuals and communities. To this end, greater scientific knowledge of climate and statistical information on the environment, economy and society will be critical.
ENDNOTES

1. Intergovernmental Panel on Climate Change (IPCC), 2007a, Climate Change 2007: Synthesis Report, Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, pp.30, 39. <back
2. CSIRO and Bureau of Meteorology, 2007, Climate Change in Australia: Technical Report 2007, pp.6, 9. <back
3. Garnaut, R., 2008, The Garnaut Climate Change Review, Cambridge University Press, p.65. <back
4. Department of Climate Change, 2009a, Australia’s National Greenhouse Accounts, National Greenhouse Gas Inventory (Accounting for the Kyoto target) May 2009, p.4. <back
5. IPCC, 2007a, op. cit., p.30. <back
6. ibid., pp.31, 72. <back
7. CSIRO and Bureau of Meteorology, op. cit., p.14. <back
8. ibid, pp.6-8. <back
9. ibid, p.32. <back
10. Fine particles suspended in the atmosphere. Main sources include: industry, vegetation burning, volcanoes and dust storms. <back
11. CSIRO, 2009, Aerosols – their part in our rainfall, <http://www.csiro.au/news/Aerosols.html>, viewed November 2009. <back
12. CSIRO and Bureau of Meteorology, op cit., p.9. <back
13. ibid. <back
14. ibid. <back
15. IPCC, 2007b, “Global Climate Projections”, in Climate Change 2007: The Physical Science Basis, Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, p.820. <back
16. IPCC, 2007c, “Summary for Policymakers”, in Climate Change 2007: The Physical Science Basis, Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, p.14. <back
17. Steffen, W., 2009, Climate Change 2009: Faster Change and More Serious Risks, Department of Climate Change, Commonwealth of Australia, p.34. <back
18. CSIRO and Bureau of Meteorology, op cit., pp.37, 72. <back
19. Garnaut, op. cit., p.153. <back
20. ibid. <back
21. Department of Climate Change, 2009a, op. cit., p.1. <back
22. ibid., pp.23–24. <back
23. International Energy Agency, 2009, op. cit., p.49. <back
24. Australian Bureau of Statistics (ABS), 2009a, Australian National Accounts, National Income, Expenditure and Product June 2009 (ABS cat. no. 5206.0 Table 02); Department of Climate Change, Australian Greenhouse Emissions Information System (AGEIS), <http://www.ageis.greenhouse.gov.au>, viewed November 2009. <back
25. Australian Greenhouse Office 2007, National Greenhouse Gas Inventory 2005 – Accounting for the 108% Target, p.5. <back
26. United Nations, 1998, Kyoto Protocol to the United Nations Framework Convention on Climate Change, p.20 (Annex B). <back
27. Department of Climate Change, 2009b, Australia’s National Greenhouse Accounts, National Inventory Report 2007 Vol. 2, p.210; Department of Climate Change, 2008, The Australian Government’s Initial Report under the Kyoto Protocol, p.2. <back
28. Department of Climate Change, 2009c, Australia’s National Greenhouse Accounts, National Inventory Report 2007 Vol. 1, p.2. <back
29. Department of Climate Change, 2009a, op. cit., p.24. <back
30. ibid., pp.5, 24. <back
31. Department of Climate Change, 2009c, op. cit., p.26. <back
32. ibid., p.27. <back
33. Department of Climate Change, 2009a, op. cit., p.16. <back
34. United Nations, 1998, op. cit., p.3. <back
35. Department of Climate Change, 2009d, National Greenhouse Accounts (NGA) Factors, p.48. <back
36. Department of Climate Change, 2009c, op. cit., p.2; Department of Climate Change, 2009a, op. cit., p.24. <back
37. Department of Climate Change, 2009a, op. cit., p.24. <back
38. Department of Climate Change, 2009c, op. cit., p.131. <back
39. Department of Climate Change, 2009a, op. cit., pp.23–24. <back
40. Department of Climate Change, 2009b, op. cit., p.190. <back
41. Garnaut, loc. cit., p.153. <back
42. ibid., p.42. <back
43. ibid., p.xiii. <back
44. ibid., pp.42–45. <back
45. ibid., pp.308–317, 356, 416–417. <back
46. ABS, 2009b, Energy Account Australia 2006–07 (ABS cat. no. 4604.0), p.41. <back
47. ibid., p.11. <back
48. ibid. <back
49. ABS, 2009c, Australian Farming in Brief 2009 (ABS cat. no. 7106.0). <back
50. Australian Government NRM team, 2008, Improving Land Management Practices, Caring for Our Country – Information Sheet, <http://www.nrm.gov.au>, viewed November 2009. <back
51. Lawson, K., Burns, K., Low, K., Heyhoe, E. and Ahammad, H., 2008, Analysing the Economic Potential of Forestry for Carbon Sequestration under Alternative Carbon Price Paths, Australian Bureau of Agricultural and Resource Economics (ABARE), Commonwealth of Australia, p.1. <back
52. Tropical Savannas CRC, 2009, The West Arnhem Land Fire Abatement Project (WALFA), <http://savanna.cdu.edu.au/information/arnhem_fire_project.html>, viewed November 2009. <back
53. Australian Bureau of Agricultural and Resource Economics (ABARE), 2009a, Energy in Australia 2009, pp.68–70. <back
54. National Transport Commission, 2009, Carbon Emissions from New Australian Vehicles Information Paper, pp.6, 15. <back
55. ibid.p.48. <back
56. Emissions not exceeding 120 g/km of CO2. <back
57. National Transport Commission, op. cit., p.28. <back
58. ABS, 2009d, Environmental Issues: Waste Management and Transport Use, Mar 2009 (ABS cat. no. 4602.0.55.002), p.59. <back
59. ibid., p.69. <back
60. ibid., p.71. <back
61. ABARE, 2009a, op. cit., p.33 and ABARE, 2009b, op. cit., p.6. <back
62. Copeland, A., 2009, Electricity Generation, Major Development Projects – April 2009 listing. Australian Bureau of Agricultural and Resource Economics (ABARE), Commonwealth of Australia. pp.5–6. <back
63. ibid. <back
64. Projects with a total capacity exceeding 30 MW. <back
65. Copeland, op. cit., p.7. <back
66. ibid., p.10. <back
67. ABS, 2008a, Environmental Issues: Energy Use and Conservation March 2008 (ABS cat. no. 4602.0.55.001), p.8. <back
68. Roarty, M. 2008. Australia’s Natural Gas: Issues and Trends, Research Paper no 25, 2007–08. Parliament of Australia, Parliamentary Library, Commonwealth of Australia, p.12. <back
69. ABS, 2008a, op. cit., p.37. <back
70. ibid., p.42. <back
71. ibid., p.53. <back
72. GreenPower 2009, National GreenPower Accreditation Program Status Report Quarter 1 – 1 January to 31 March 2009, p.2; GreenPower, 2005, National Green Power Accreditation Program Quarterly Status Report, 1 January to 31 March 2005, p.1. <back
73. ABS, 2008a, op. cit. p.96. <back
74. ibid., p.97. <back
75. ABS, 2009e, Environmental Views and Behaviour 2007–08 (2nd issue) (ABS cat. no. 4626.0.55.001), Tables 09 and 11. <back
76. Milne, G. and Reidy, C., 2008, “6.1 Energy Use”, in Your Home Technical Manual, Edition 4, a joint initiative of the Australian Government and the design and construction Industries, <http://www.yourhome.gov.au/technical/fs61.html> viewed September 2009. <back
77. ABS, 2008a, op. cit., p.67. <back
78. ABS, 2009e, op. cit., Tables 25 and 27. <back
79. ABS, 2008a, op. cit., p.7. <back
80. Stern, N., 2006, Stern Review on the Economics of Climate Change (Executive Summary p.i). <back
81. Garnaut, op. cit., p.300. <back
82. Carbon dioxide emissions. <back
83. Garnaut, op. cit., pp.300–309. <back
84. ibid., p.388. <back
85. ABS, 2007a, Australian Social Trends 2007 (ABS cat. no. 4102.0), p.168. <back
86. Garnaut, op. cit. pp.387–389. <back
87. ibid., p.363. <back
88. ibid. <back
89. ABS, 2008b, Water and the Murray-Darling Basin – A Statistical Profile 2000–01 to 2005–06 (ABS cat. no. 4610.0.55.007), p.53. <back
90. Jones, R.N., and Preston, B.L.,2006, Climate Change Impacts, Risk and the Benefits of Mitigation, A Report for the Energy Futures Program, CSIRO, pp.45–46, <back
91. CSIRO and Bureau of Meteorology, op. cit., p.83. <back
92. ABS, 2006b, Water Account, Australia 2004–05 (ABS cat. no. 4610.0), pp.115-125. <back
93. ibid., p.124. <back
94. ibid. <back
95. Murray-Darling Basin Authority, 2009, River Murray System Drought Update (Issue 21: November 2009), <http://www.mdba.gov.au/system/files/drought-update-November-2009.pdf>, viewed November 2009. <back
96. ABS, 2007b, Natural Resource Management on Australian Farms 2004–05 (ABS cat. no. 4620.0), pp.10, 19. <back
97. ibid., p.25. <back
98. ibid. <back
99. Excess nutrients in a water body, often leading to enhanced plant growth and degradation of water/habitat quality. <back
100. Biodiversity and Climate Change Expert Advisory Group, 2009, Australia’s Biodiversity and Climate Change: Technical Synthesis, Australian Government Department of Climate Change, Canberra, p.31. <back
101. National Water Commission, 2009, Australian Water Reform 2009, p.62. <back
102. ABS, 2006b, op. cit., pp.54–55. <back
103. National Water Commission, 2009, op. cit., p.65. <back
104. ABS, 2006b, op. cit., p.103. <back
105. ABS, 2007c, Environmental Issues: People’s Views and Practices March 2007 (ABS cat. no. 4602.0), p.60. <back
106. ibid., p.19. <back
107. ibid. <back
108. ABS, 2009f, Value of Agricultural Commodities Produced, Australia 2007–08 (ABS cat. no. 7503.0), p.3. <back
109. ABS, 2009g, Labour Force, Australia, Detailed – Electronic Delivery (ABS cat. no. 6291.0.55.003), Table 06. <back
110. Australian Bureau of Agricultural and Resource Economics, 2008, Australian Commodity Statistics 2008., p.4. <back
111. Garnaut, op. cit. p.129. <back
112. ABS, 2008c, Farm Management and Climate 2006–07 (ABS cat. no. 4625.0), p.11. <back
113. ibid., p.8. <back
114. ibid. <back
115. ABS, 2009h, Water Use on Australian Farms 2007-08 (ABS cat. no. 4618.0), p.4. <back
116. ABS, 2009i, Experimental Estimates of the Gross Value of Irrigated Agricultural Production, 2000–01 to 2006–07 (cat. no. 4610.0.55.008). <back
117. ibid. <back
118. ABS, 2008b, op. cit., p.1. <back
119. ibid., p.3. <back
120. ibid., p.1. <back
121. ibid., p.2. <back
122. ibid., p.3. <back
123. Garnaut, op. cit. p.130. <back
124. ibid., p.377. <back
125. ABS, 2008b op cit., p.81. <back
126. ibid. <back
127. Steffen, W., Burbidge, A., Hughes, L., Kitching, R., Lindenmayer, D., Musgrave, W., Stafford Smith, M. and Werner, P., 2009, Australia’s Biodiversity and Climate Change: Technical Synthesis, Prepared for the Australian Government by the Biodiversity and Climate Change Expert Group, Department of Climate Change, p.19. <back
128. ibid., p.21. <back
129. Garnaut, op. cit., p.380. <back
130. Tourism Research Australia, 2009, Nature Tourism in Australia 2008, p.1. <back
131. ibid. <back
132. ibid. <back
133. Garnaut, op. cit., pp.133–134. <back
134. ibid. <back
135. House of Representatives Standing Committee on Climate Change, Water, Environment and the Arts, 2009, Managing Our Coastal Zone in a Changing Climate, p.1. <back
136. Hennessey, K., Fitzharris, B., Bates, B.C., Harvey, N., Howden, S.M., Hughes, L., Salinger, J. and Warrick, R., 2007, “Australia and New Zealand”, in Climate Change 2007: Impacts, Adaptation and Vulnerability, Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, p.520. <back
137. ibid. <back
138. ibid. <back
139. ABS, 2009j, Regional Population Growth, Australia 2007–08 (ABS cat. no. 3218.0). <back
140. World Health Organisation, 2009, Climate Change and Human Health, <http://www.who.int/globalchange/en/>, viewed September 2009. <back
141. Garnaut, op. cit., p.139. <back
142. World Health Organisation, 2009, 10 Facts on Climate Change and Health, <http://www.who.int/features/factfiles/climate_change/en/index.html> viewed September 2009. <back
143. Garnaut, loc. cit., p.139; PMSEIC Independent Working Group, 2007, Climate Change in Australia: Regional Impacts and Adaptation – Managing the Risk for Australia, Report prepared for the Prime Minister’s Science, Engineering and Innovation Council, p.28. <back
144. PMSEIC Independent Working Group, 2007, Climate Change in Australia: Regional Impacts and Adaptation – Managing the Risk for Australia, Report prepared for the Prime Minister’s Science, Engineering and Innovation Council, p.28. <back
145. Garnaut, op. cit., p.117. <back
146. IPCC, 2007a, op. cit., p. 53. <back
147. ibid., pp.8–9. <back
148. ABS, 2009k, Population by Age and Sex, Australian States and Territories June 2009 (ABS cat. no. 3201.0). <back
149. ABS, 2008d, Population Projections Australia 2006 to 2101 (ABS cat. no. 3222.0), p.44. <back
150. ABS, 2009e, op. cit., Table 07. <back