Wednesday 30 May 2012

Message of the United Nations Secretary-General, Ban Ki-Moon - 5 June 2012


Message of the United Nations Secretary-General, Ban Ki-Moon - 5 June 2012

As the world gears up for the United Nations Conference on Sustainable Development (Rio+20), World Environment Day is an opportunity to highlight the need for a paradigm shift towards a more sustainable world.  This year’s theme, “Green Economy: Does it include you?”, underscores the need for everyone to play their part in keeping humankind's ecological footprint within planetary boundaries.
The world’s population stands at 7 billion and may rise to more than 9 billion by 2050.  This means greater pressure on already crowded cities – where more than half of all people now live – and on natural resources, as demand for food, water and energy rises.  It also means more people in search of decent jobs.  Globally, 1.3 billion people are currently unemployed or under-employed.  An estimated half billion more will join the job market over the next decade. 
Sustainability entails providing opportunity for all by balancing the social, economic and environmental dimensions of development.  We have to rebut the myth that there is conflict between economic and environmental health.  With smart policies and the right investments, countries can protect their environment, grow their economies, generate decent jobs and accelerate social progress.
Rio+20 is our opportunity to deepen global commitment to sustainable development.  In Rio, we should agree that measuring growth and wealth by Gross Domestic Product alone is inadequate.  We should agree that the world needs a set of sustainable development goals that will build on the Millennium Development Goals.  And we should make progress on some of the building blocks of sustainability – energy, water, food, cities, oceans, jobs and the empowerment of women. 
Sustainability is gaining prominence on the public policy agenda in both developed and developing nations.  The UN itself is working towards climate neutrality and sustainable management of our offices and activities.  In Rio, we must mobilize the partnerships we need to shift the world onto a more sustainable trajectory of growth and development.  On this World Environment Day, in advance of this historic conference, I urge governments, businesses and all members of society to make the holistic choices that will ensure a sustainable future – the future we want. 
WED home page

INDIA 2012 MONSOON ARRIVAL DATES

‘Climate killed Harappan civilization'

WASHINGTON: Climate change may be the main culprit behind the collapse of the Indus Valley Civilization around 4,000 years ago, says a new study, which also claims to have resolved the long-standing debate over the source and fate of the Saraswati, a sacred river in Hindu mythology.
The study, combining the latest archaeological data along with state-of-the-art geoscience technologies, suggested that decline in monsoon rains led to weakened river dynamics, and played a critical role both in the development and the fall of the Harappan culture, which relied on river floods to fuel their agricultural surpluses. The international team, which published their findings in the journal Proceedings of the National Academy of Sciences , used satellite photos and topographic data to make and analyse digital maps of landforms constructed by the Indus and other neighbouring rivers, which were then probed in the field by drilling, coring, and even manuallydug trenches. Collected samples were used to determine the sediments' origins, whether brought in by rivers or wind, and their age, in order to develop a chronology of landscape changes.
 "We reconstructed the dynamic landscape of the plain where the Indus civilization developed 5,200 years ago, built its cities, and disintegrated between 3,900 and 3,000 years ago," said lead researcher Liviu Giosan, a geologist with Woods Hole Oceanographic Institution in the US. "Our study suggests that the decline in monsoon rains led to weakened river dynamics , and played a key role both in development and the fall of Harappan culture," he said.
The research, which was conducted between 2003 and 2008, also claimed that the mythical Saraswati river was actually not fed by glaciers in the Himalayas as believed. Rather, it was a perennial monsoon-supported watercourse and aridification reduced it to short seasonal flows, the researchers said.

Greenhouse gas

Greenhouse gas

From Wikipedia, the free encyclopedia


Simple diagram of greenhouse effect.
A greenhouse gas (sometimes abbreviated GHG) is a gas in an atmosphere that absorbs and emits radiation within the thermal infrared range. This process is the fundamental cause of the greenhouse effect.[1] The primary greenhouse gases in the Earth's atmosphere are water vapour, carbon dioxide, methane, nitrous oxide, and ozone. In the Solar System, the atmospheres of Venus, Mars, and Titan also contain gases that cause greenhouse effects. Greenhouse gases greatly affect the temperature of the Earth; without them, Earth's surface would be on average about 33 °C (59 °F)[note 1] colder than at present.[2][3][4]
However, since the beginning of the Industrial Revolution, the burning of fossil fuels has contributed to the increase in carbon dioxide in the atmosphere from 280 ppm to 390 ppm, despite the uptake of a large portion of the emissions through various natural "sinks" involved in the carbon cycle.[5][6] Anthropogenic carbon dioxide (CO2 ) emissions (i.e., emissions produced by human activities) come from combustion of carbonaceous fuels, principally wood, coal, oil, and natural gas.[7]

Gases in Earth's atmosphere

Greenhouse gases


Atmospheric absorption and scattering at different electromagnetic wavelengths. The largest absorption band of carbon dioxide is in the infrared.
Greenhouse gases are those that can absorb and emit infrared radiation.[1] In order, the most abundant greenhouse gases in Earth's atmosphere are:
Atmospheric concentrations of greenhouse gases are determined by the balance between sources (emissions of the gas from human activities and natural systems) and sinks (the removal of the gas from the atmosphere by conversion to a different chemical compound).[8] The proportion of an emission remaining in the atmosphere after a specified time is the "Airborne fraction" (AF). More precisely, the annual AF is the ratio of the atmospheric increase in a given year to that year’s total emissions. For CO2 the AF over the last 50 years (1956–2006) has been increasing at 0.25 ± 0.21%/year.[9]

Non-greenhouse gases

Although contributing to many other physical and chemical reactions, the major atmospheric constituents, nitrogen (N2), oxygen (O2), and argon (Ar), are not greenhouse gases. This is because molecules containing two atoms of the same element such as N2 and O2 and monatomic molecules such as Argon (Ar) have no net change in their dipole moment when they vibrate and hence are almost totally unaffected by infrared light. Although molecules containing two atoms of different elements such as carbon monoxide (CO) or hydrogen chloride (HCl) absorb IR, these molecules are short-lived in the atmosphere owing to their reactivity and solubility. Because they do not contribute significantly to the greenhouse effect, they are usually omitted when discussing greenhouse gases.

Non-greenhouse gases that have an indirect radiative effect


MOPITT 2000 global carbon monoxide.
Carbon monoxide has an indirect radiative effect by elevating concentrations of methane and tropospheric ozone through scavenging of atmospheric constituents (e.g., the hydroxyl radical, OH) that would otherwise destroy them. Carbon monoxide is created when carbon-containing fuels are burned incompletely. Through natural processes in the atmosphere, it is eventually oxidized to carbon dioxide. Carbon monoxide has an atmospheric lifetime of only a few months[10] and as a consequence is spatially more variable than longer-lived gases.

Contribution of clouds to Earth's greenhouse effect

The major non-gas contributor to the Earth's greenhouse effect, clouds, also absorb and emit infrared radiation and thus have an effect on radiative properties of the greenhouse gases. Clouds are water droplets or ice crystals suspended in the atmosphere.[11][12]

Impact of a given gas on the overall greenhouse effect

The contribution of each gas to the greenhouse effect is affected by the characteristics of that gas, its abundance, and any indirect effects it may cause. For example, on a molecule-for-molecule basis the direct radiative effects of methane is about 72 times stronger than carbon dioxide over a 20 year time frame[13] but it is present in much smaller concentrations so that its total direct radiative effect is smaller, and it has a shorter atmospheric lifetime. On the other hand, in addition to its direct radiative impact methane has a large indirect radiative effect because it contributes to ozone formation. Shindell et al. (2005)[14] argue that the contribution to climate change from methane is at least double previous estimates as a result of this effect.[15]
When these gases are ranked by their direct contribution to the greenhouse effect, the most important are:[11]
Gas
Formula
Contribution
(%)
Water vapor H2O 36 – 72 %  
Carbon dioxide CO2 9 – 26 %
Methane CH4 4 – 9 %  
Ozone O3 3 – 7 %  
In addition to the main greenhouse gases listed above, other greenhouse gases include sulfur hexafluoride, hydrofluorocarbons and perfluorocarbons (see IPCC list of greenhouse gases). Some greenhouse gases are not often listed. For example, nitrogen trifluoride has a high global warming potential (GWP) but is only present in very small quantities.[16]

Proportion of direct effects at a given moment

It is not possible to state that a certain gas causes an exact percentage of the greenhouse effect. This is because some of the gases absorb and emit radiation at the same frequencies as others, so that the total greenhouse effect is not simply the sum of the influence of each gas. The higher ends of the ranges quoted are for each gas alone; the lower ends account for overlaps with the other gases.[11][12] In addition, some gases such as methane are known to have large indirect effects that are still being quantified.[17]

Atmospheric lifetime

Aside from water vapor, which has a residence time of about nine days,[18] major greenhouse gases are well-mixed, and take many years to leave the atmosphere.[19] Although it is not easy to know with precision how long it takes greenhouse gases to leave the atmosphere, there are estimates for the principal greenhouse gases. Jacob (1999)[20] defines the lifetime \tau of an atmospheric species X in a one-box model as the average time that a molecule of X remains in the box. Mathematically \tau can be defined as the ratio of the mass m (in kg) of X in the box to its removal rate, which is the sum of the flow of X out of the box (F_{out}), chemical loss of X (L), and deposition of X (D) (all in kg/sec): \tau = \frac{m}{F_{out}+L+D}.[20] If one stopped pouring any of this gas into the box, then after a time \tau, its concentration would be about halved.
The atmospheric lifetime of a species therefore measures the time required to restore equilibrium following a sudden increase or decrease in its concentration in the atmosphere. Individual atoms or molecules may be lost or deposited to sinks such as the soil, the oceans and other waters, or vegetation and other biological systems, reducing the excess to background concentrations. The average time taken to achieve this is the mean lifetime.
Carbon dioxide has a variable atmospheric lifetime, and cannot be specified precisely.[21] The atmospheric lifetime of CO2 is estimated of the order of 30–95 years.[22] This figure accounts for CO2 molecules being removed from the atmosphere by mixing into the ocean, photosynthesis, and a few other processes. However, this excludes the balancing fluxes of CO2 into the atmosphere from the geological reservoirs, which have slower characteristic rates.[23] While more than half of the CO2 emitted is currently removed from the atmosphere within a century, some fraction (about 20%) of emitted CO2 remains in the atmosphere for many thousands of years.[24][25][26]

Global warming potential

The global warming potential (GWP) depends on both the efficiency of the molecule as a greenhouse gas and its atmospheric lifetime. GWP is measured relative to the same mass of CO2 and evaluated for a specific timescale. Thus, if a gas has a high radiative forcing but also a short lifetime, it will have a large GWP on a 20 year scale but a small one on a 100 year scale. Conversely, if a molecule has a longer atmospheric lifetime than CO2 its GWP will increase with the timescale considered. Carbon dioxide is defined to have a GWP of 1 over all time periods.
Methane has an atmospheric lifetime of 12 ± 3 years and a GWP of 72 over 20 years, 25 over 100 years and 7.6 over 500 years. The decrease in GWP at longer times is because methane is degraded to water and CO2 through chemical reactions in the atmosphere.
Examples of the atmospheric lifetime and GWP relative to CO2 for several greenhouse gases are given in the following table:[13]
Atmospheric lifetime and GWP relative to CO2 at different time horizon for various greenhouse gases.
Gas name Chemical
formula
Lifetime
(years)
Global warming potential (GWP) for given time horizon
20-yr 100-yr 500-yr
Carbon dioxide CO2 See above 1 1 1
Methane CH4 12 72 25 7.6
Nitrous oxide N2O 114 289 298 153
CFC-12 CCl2F2 100 11 000 10 900 5 200
HCFC-22 CHClF2 12 5 160 1 810 549
Tetrafluoromethane CF4 50 000 5 210 7 390 11 200
Hexafluoroethane C2F6 10 000 8 630 12 200 18 200
Sulphur hexafluoride SF6 3 200 16 300 22 800 32 600
Nitrogen trifluoride NF3 740 12 300 17 200 20 700
The use of CFC-12 (except some essential uses) has been phased out due to its ozone depleting properties.[27] The phasing-out of less active HCFC-compounds will be completed in 2030.[28]

Natural and anthropogenic sources


400,000 years of ice core data.

Top: Increasing atmospheric carbon dioxide levels as measured in the atmosphere and reflected in ice cores. Bottom: The amount of net carbon increase in the atmosphere, compared to carbon emissions from burning fossil fuel.
Aside from purely human-produced synthetic halocarbons, most greenhouse gases have both natural and human-caused sources. During the pre-industrial Holocene, concentrations of existing gases were roughly constant. In the industrial era, human activities have added greenhouse gases to the atmosphere, mainly through the burning of fossil fuels and clearing of forests.[29][30]
The 2007 Fourth Assessment Report compiled by the IPCC (AR4) noted that "changes in atmospheric concentrations of greenhouse gases and aerosols, land cover and solar radiation alter the energy balance of the climate system", and concluded that "increases in anthropogenic greenhouse gas concentrations is very likely to have caused most of the increases in global average temperatures since the mid-20th century".[31] In AR4, "most of" is defined as more than 50%.
Gas Preindustrial level Current level   Increase since 1750   Radiative forcing (W/m2)
Carbon dioxide 280 ppm  396 ppm 116 ppm 1.46
Methane 700 ppb 1745 ppb 1045 ppb  0.48
Nitrous oxide 270 ppb  314 ppb  44 ppb 0.15
CFC-12 0  533 ppt 533 ppt 0.17
Ice cores provide evidence for variation in greenhouse gas concentrations over the past 800,000 years. Both CO2 and CH4 vary between glacial and interglacial phases, and concentrations of these gases correlate strongly with temperature. Direct data does not exist for periods earlier than those represented in the ice core record, a record which indicates CO2 mole fractions staying within a range of between 180 ppm and 280 ppm throughout the last 800,000 years, until the increase of the last 250 years. However, various proxies and modeling suggests larger variations in past epochs; 500 million years ago CO2 levels were likely 10 times higher than now.[32] Indeed higher CO2 concentrations are thought to have prevailed throughout most of the Phanerozoic eon, with concentrations four to six times current concentrations during the Mesozoic era, and ten to fifteen times current concentrations during the early Palaeozoic era until the middle of the Devonian period, about 400 Ma.[33][34][35] The spread of land plants is thought to have reduced CO2 concentrations during the late Devonian, and plant activities as both sources and sinks of CO2 have since been important in providing stabilising feedbacks.[36] Earlier still, a 200-million year period of intermittent, widespread glaciation extending close to the equator (Snowball Earth) appears to have been ended suddenly, about 550 Ma, by a colossal volcanic outgassing which raised the CO2 concentration of the atmosphere abruptly to 12%, about 350 times modern levels, causing extreme greenhouse conditions and carbonate deposition as limestone at the rate of about 1 mm per day.[37] This episode marked the close of the Precambrian eon, and was succeeded by the generally warmer conditions of the Phanerozoic, during which multicellular animal and plant life evolved. No volcanic carbon dioxide emission of comparable scale has occurred since. In the modern era, emissions to the atmosphere from volcanoes are only about 1% of emissions from human sources.[37][38]

Anthropogenic greenhouse gases


Modern global anthropogenic carbon emissions.

Global anthropogenic greenhouse gas emissions broken down into 8 different sectors for the year 2000.

Per capita anthropogenic greenhouse gas emissions by country for the year 2000 including land-use change.
Since about 1750 human activity has increased the concentration of carbon dioxide and other greenhouse gases. Measured atmospheric concentrations of carbon dioxide are currently 100 ppm higher than pre-industrial levels.[39] Natural sources of carbon dioxide are more than 20 times greater than sources due to human activity,[40] but over periods longer than a few years natural sources are closely balanced by natural sinks, mainly photosynthesis of carbon compounds by plants and marine plankton. As a result of this balance, the atmospheric mole fraction of carbon dioxide remained between 260 and 280 parts per million for the 10,000 years between the end of the last glacial maximum and the start of the industrial era.[41]
It is likely that anthropogenic warming, such as that due to elevated greenhouse gas levels, has had a discernible influence on many physical and biological systems. Warming is projected to affect various issues such as freshwater resources, industry, food and health.[42]
The main sources of greenhouse gases due to human activity are:
  • burning of fossil fuels and deforestation leading to higher carbon dioxide concentrations in the air. Land use change (mainly deforestation in the tropics) account for up to one third of total anthropogenic CO2 emissions.[41]
  • livestock enteric fermentation and manure management,[43] paddy rice farming, land use and wetland changes, pipeline losses, and covered vented landfill emissions leading to higher methane atmospheric concentrations. Many of the newer style fully vented septic systems that enhance and target the fermentation process also are sources of atmospheric methane.
  • use of chlorofluorocarbons (CFCs) in refrigeration systems, and use of CFCs and halons in fire suppression systems and manufacturing processes.
  • agricultural activities, including the use of fertilizers, that lead to higher nitrous oxide (N2O) concentrations.
The seven sources of CO2 from fossil fuel combustion are (with percentage contributions for 2000–2004):[44]
Seven main fossil fuel
combustion sources
Contribution
(%)
Liquid fuels (e.g., gasoline, fuel oil) 36 %
Solid fuels (e.g., coal) 35 %
Gaseous fuels (e.g., natural gas) 20 %
Cement production  3 %
Flaring gas industrially and at wells < 1 %  
Non-fuel hydrocarbons < 1 %  
"International bunker fuels" of transport
not included in national inventories
 4 %
The US Environmental Protection Agency (EPA) ranks the major greenhouse gas contributing end-user sectors in the following order: industrial, transportation, residential, commercial and agricultural.[45] Major sources of an individual's greenhouse gas include home heating and cooling, electricity consumption, and transportation. Corresponding conservation measures are improving home building insulation, installing geothermal heat pumps and compact fluorescent lamps, and choosing energy-efficient vehicles.
Carbon dioxide, methane, nitrous oxide and three groups of fluorinated gases (sulfur hexafluoride, HFCs, and PFCs) are the major greenhouse gases and the subject of the Kyoto Protocol, which came into force in 2005.[46]
Although CFCs are greenhouse gases, they are regulated by the Montreal Protocol, which was motivated by CFCs' contribution to ozone depletion rather than by their contribution to global warming. Note that ozone depletion has only a minor role in greenhouse warming though the two processes often are confused in the media.
On December 7, 2009, the US Environmental Protection Agency released its final findings on greenhouse gases, declaring that "greenhouse gases (GHGs) threaten the public health and welfare of the American people". The finding applied to the same "six key well-mixed greenhouse gases" named in the Kyoto Protocol: carbon dioxide, methane, nitrous oxide, hydrofluorocarbons, perfluorocarbons, and sulfur hexafluoride.[47][48]

Role of water vapor


Increasing water vapor in the stratosphere at Boulder, Colorado.
Water vapor accounts for the largest percentage of the greenhouse effect, between 36% and 66% for clear sky conditions and between 66% and 85% when including clouds.[12] Water vapor concentrations fluctuate regionally, but human activity does not significantly affect water vapor concentrations except at local scales, such as near irrigated fields. The atmospheric concentration of vapor is highly variable, from less than 0.01% in extremely cold regions up to 20% in warm, humid regions.[49]
The average residence time of a water molecule in the atmosphere is only about nine days, compared to years or centuries for other greenhouse gases such as CH4 and CO2. Thus, water vapor responds to and amplifies effects of the other greenhouse gases. The Clausius-Clapeyron relation establishes that air can hold more water vapor per unit volume when it warms. This and other basic principles indicate that warming associated with increased concentrations of the other greenhouse gases also will increase the concentration of water vapor. Because water vapor is a greenhouse gas, this results in further warming and so is a "positive feedback" that amplifies the original warming. Eventually other earth processes offset these positive feedbacks, stabilizing the global temperature at a new equilibrium and preventing the loss of Earth's water through a Venus-like runaway greenhouse effect.[50]

Atmospheric concentration

Measurements from Antarctic ice cores show that before industrial emissions started atmospheric CO2 mole fractions were about 280 parts per million (ppm), and stayed between 260 and 280 during the preceding ten thousand years.[51] Carbon dioxide mole fractions in the atmosphere have gone up by approximately 35 percent since the 1900s, rising from 280 parts per million by volume to 387 parts per million in 2009. One study using evidence from stomata of fossilized leaves suggests greater variability, with carbon dioxide mole fractions above 300 ppm during the period seven to ten thousand years ago,[52] though others have argued that these findings more likely reflect calibration or contamination problems rather than actual CO2 variability.[53][54] Because of the way air is trapped in ice (pores in the ice close off slowly to form bubbles deep within the firn) and the time period represented in each ice sample analyzed, these figures represent averages of atmospheric concentrations of up to a few centuries rather than annual or decadal levels.
Since the beginning of the Industrial Revolution, the concentrations of most of the greenhouse gases have increased. For example, the mole fraction of carbon dioxide has increased from 280 ppm by about 36% to 380 ppm, or 100 ppm over modern pre-industrial levels. The first 50 ppm increase took place in about 200 years, from the start of the Industrial Revolution to around 1973; however the next 50 ppm increase took place in about 33 years, from 1973 to 2006.[55]
Recent data also shows that the concentration is increasing at a higher rate. In the 1960s, the average annual increase was only 37% of what it was in 2000 through 2007.[56]
Today, the stock of carbon in the atmosphere increases by more than 3 million tonnes per annum (0.04%) compared with the existing stock.[clarification needed] This increase is the result of human activities by burning fossil fuels, deforestation and forest degradation in tropical and boreal regions.[57]
The other greenhouse gases produced from human activity show similar increases in both amount and rate of increase. Many observations are available online in a variety of Atmospheric Chemistry Observational Databases.
Relevant to radiative forcing
Gas Current (1998)
Amount by volume
Increase
(absolute, ppm)
over pre-industrial (1750)
Increase
(relative, %)
over pre-industrial (1750)
Radiative
forcing
(W/m2)
Carbon dioxide  365 ppm
(383 ppm, 2007.01)
   87 ppm
(105 ppm, 2007.01)
31 %
(38 %, 2007.01)
1.46
(~1.53, 2007.01)
Methane 1745 ppb 1045 ppb 150 % 0.48
Nitrous oxide  314 ppb    44 ppb 16 % 0.15
Relevant to both radiative forcing and ozone depletion; all of the following have no natural sources and hence zero amounts pre-industrial
Gas Current (1998)
Amount by volume
Radiative forcing
(W/m2)
CFC-11 268 ppt 0.07
CFC-12 533 ppt 0.17
CFC-113  84 ppt 0.03
Carbon tetrachloride 102 ppt 0.01
HCFC-22  69 ppt 0.03
(Source: IPCC radiative forcing report 1994 updated (to 1998) by IPCC TAR table 6.1[58][59] ).

Greenhouse gas emissions ("sources")


Recent year-to-year increase of atmospheric CO2.
Between the period 1970 to 2004, GHG emissions (measured in CO2-equivalent)[60] increased at an average rate of 1.6% per year, with CO2 emissions from the use of fossil fuels growing at a rate of 1.9% per year.[61][62] Total anthropogenic emissions at the end of 2009 were estimated at 49.5 gigatonnes CO2-equivalent.[63] These emissions include CO2 from fossil fuel use and from land use, as well as emissions of methane, nitrous oxide and other GHGs covered by the Kyoto Protocol.
At present, the two primary sources of CO2 emissions are from burning coal used for electricity generation and petroleum used for motor transport.[citation needed]

Regional and national attribution of emissions


Major greenhouse gas trends.
There are several different ways of measuring GHG emissions (see World Bank (2010, p. 362) for a table of national emissions data).[64]
Some variables that have been reported[65] include:
  • Definition of measurement boundaries. Emissions can be attributed geographically, to the area where they were emitted (the territory principle) or by the activity principle to the territory that caused the emissions to be produced. These two principles would result in different totals when measuring for example the importation of electricity from one country to another or the emissions at an international airport.
  • The time horizon of different GHGs. Contribution of a given GHG is reported as a CO2 equivalent; the calculation to determine this takes into account how long that gas remains in the atmosphere. This is not always known accurately and calculations must be regularly updated to take into account new information.
  • What sectors are included in the calculation (e.g. energy industries, industrial processes, agriculture etc.). There is often a conflict between transparency and availability of data.
  • The measurement protocol itself. This may be via direct measurement or estimation; the four main methods are the emission factor-based method, the mass balance method, the predictive emissions monitoring system and the continuing emissions monitoring systems. The methods differ in accuracy, but also in cost and usability.
The different measures are sometimes used by different countries in asserting various policy/ethical positions to do with climate change (Banuri et al., 1996, p. 94).[66] This use of different measures leads to a lack of comparability, which is problematic when monitoring progress towards targets. There are arguments for the adoption of a common measurement tool, or at least the development of communication between different tools.[65]
Emissions may be measured over long time periods. This measurement type is called historical or cumulative emissions. Cumulative emissions give some indication of who is responsible for the build-up in the atmospheric concentration of GHGs (IEA, 2007, p. 199).[67]
The national accounts balance would be positively related to carbon emissions. The national accounts balance shows the difference between exports and imports. For many richer nations, such as the United States, the accounts balance is negative because more goods are imported than they are exported. This is mostly due to the fact that it is cheaper to produce goods outside of developed countries, leading the economies of developed countries to become increasingly dependent on services and not goods. We believed that a positive accounts balance would means that more production was occurring in a country, so more factories working would increase carbon emission levels.(Holtz-Eakin, 1995, pp.;85;101).[68]
Emissions may also be measured across shorter time periods. Emissions changes may, for example, be measured against a base year of 1990. 1990 was used in the United Nations Framework Convention on Climate Change (UNFCCC) as the base year for emissions, and is also used in the Kyoto Protocol (some gases are also measured from the year 1995) (Grubb, 2003, pp. 146, 149).[69] A country's emissions may also be reported as a proportion of global emissions for a particular year.
Another measurement is of per capita emissions. This divides a country's total annual emissions by its mid-year population (World Bank, 2010, p. 370). Per capita emissions may be based on historical or annual emissions (Banuri et al., 1996, pp. 106–107).

Greenhouse gas intensity and land-use change

Greenhouse gas intensity in the year 2000, including land-use change.
Cumulative energy-related CO2 emissions between the years 1850-2005 grouped into low-income, middle-income, high-income, the EU-15, and the OECD countries.
Cumulative energy-related CO2 emissions between the years 1850-2005 for individual countries.
Map of cumulative per capita anthropogenic atmospheric CO2 emissions by country. Cumulative emissions include land use change, and are measured between the years 1950 and 2000.
Regional trends in annual CO2 emissions from fuel combustion between 1971 and 2009.
Regional trends in annual per capita CO2 emissions from fuel combustion between 1971 and 2009.
The first figure shown opposite is based on data from the World Resources Institute, and shows a measurement of GHG emissions for the year 2000 according to greenhouse gas intensity and land-use change. Herzog et al. (2006, p. 3) defined greenhouse gas intensity as GHG emissions divided by economic output.[70] GHG intensities are subject to uncertainty over whether they are calculated using market exchange rates (MER) or purchasing power parity (PPP) (Banuri et al., 1996, p. 96).[66] Calculations based on MER suggest large differences in intensities between developed and developing countries, whereas calculations based on PPP show smaller differences.
Land-use change, e.g., the clearing of forests for agricultural use, can affect the concentration of GHGs in the atmosphere by altering how much carbon flows out of the atmosphere into carbon sinks.[71] Accounting for land-use change can be understood as an attempt to measure “net” emissions, i.e., gross emissions from all GHG sources minus the removal of emissions from the atmosphere by carbon sinks (Banuri et al., 1996, pp. 92–93).
There are substantial uncertainties in the measurement of net carbon emissions.[72] Additionally, there is controversy over how carbon sinks should be allocated between different regions and over time (Banuri et al., 1996, p. 93). For instance, concentrating on more recent changes in carbon sinks is likely to favour those regions that have deforested earlier, e.g., Europe.

Cumulative and historical emissions

Cumulative anthropogenic (i.e., human-emitted) emissions of CO2 from fossil fuel use are a major cause of global warming,[73] and give some indication of which countries have contributed most to human-induced climate change.[74]:15
Top-5 historic CO2 contributors by region over the years 1800 to 1988 (in %)
Region Industrial
CO2
Total
CO2
OECD North America 33.2 29.7
OECD Europe 26.1 16.6
Former USSR 14.1 12.5
China   5.5   6.0
Eastern Europe   5.5   4.8
The table above to the left is based on Banuri et al. (1996, p. 94).[66] Overall, developed countries accounted for 83.8% of industrial CO2 emissions over this time period, and 67.8% of total CO2 emissions. Developing countries accounted for industrial CO2 emissions of 16.2% over this time period, and 32.2% of total CO2 emissions. The estimate of total CO2 emissions includes biotic carbon emissions, mainly from deforestation. Banuri et al. (1996, p. 94) calculated per capita cumulative emissions based on then-current population. The ratio in per capita emissions between industrialized countries and developing countries was estimated to be more than 10 to 1.
Including biotic emissions brings about the same controversy mentioned earlier regarding carbon sinks and land-use change (Banuri et al., 1996, pp. 93–94). The actual calculation of net emissions is very complex, and is affected by how carbon sinks are allocated between regions and the dynamics of the climate system.
Non-OECD countries accounted for 42% of cumulative energy-related CO2 emissions between 1890-2007.[75]:179-180 Over this time period, the US accounted for 28% of emissions; the EU, 23%; Russia, 11%; China, 9%; other OECD countries, 5%; Japan, 4%; India, 3%; and the rest of the world, 18%.[75]:179-180

Changes since a particular base year

Between 1970-2004, global growth in annual CO2 emissions was driven by North America, Asia, and the Middle East.[76] The sharp acceleration in CO2 emissions since 2000 to more than a 3% increase per year (more than 2 ppm per year) from 1.1% per year during the 1990s is attributable to the lapse of formerly declining trends in carbon intensity of both developing and developed nations. China was responsible for most of global growth in emissions during this period. Localised plummeting emissions associated with the collapse of the Soviet Union have been followed by slow emissions growth in this region due to more efficient energy use, made necessary by the increasing proportion of it that is exported.[44] In comparison, methane has not increased appreciably, and N2O by 0.25% y−1.
Using different base years for measuring emissions has an effect on estimates of national contributions to global warming.[74]:17-18[77] This can be calculated by dividing a country's highest contribution to global warming starting from a particular base year, by that country's minimum contribution to global warming starting from a particular base year. Choosing between different base years of 1750, 1900, 1950, and 1990 has a significant effect for most countries.[74]:17-18 Within the G8 group of countries, it is most significant for the UK, France and Germany. These countries have a long history of CO2 emissions (see the section on Cumulative and historical emissions).

Annual emissions

Annual per capita emissions in the industrialized countries are typically as much as ten times the average in developing countries.[69]:144 Due to China's fast economic development, its annual per capita emissions are quickly approaching the levels of those in the Annex I group of the Kyoto Protocol (i.e., the developed countries excluding the USA).[78] Other countries with fast growing emissions are South Korea, Iran, and Australia. On the other hand, annual per capita emissions of the EU-15 and the USA are gradually decreasing over time.[78] Emissions in Russia and the Ukraine have decreased fastest since 1990 due to economic restructuring in these countries.[79]
Energy statistics for fast growing economies are less accurate than those for the industrialized countries. For China's annual emissions in 2008, the Netherlands Environmental Assessment Agency estimated an uncertainty range of about 10%.[78]

Top emitters

Bar graph of annual per capita CO2 emissions from fuel combustion for 140 countries in 2009.
Bar graph of cumulative energy-related per capita CO2 emissions between 1850-2008 for 185 countries.

Annual

In 2009, the annual top ten emitting countries accounted for about two-thirds of the world's annual energy-related CO2 emissions.[80]
Top-10 annual energy-related CO2 emitters for the year 2009[81]
Country  % of global total
annual emissions
Tonnes of GHG
per capita
People's Rep. of China 23.6 5.13
United States 17.9 16.9
India 5.5 1.37
Russian Federation 5.3 10.8
Japan 3.8 8.6
Germany 2.6 9.2
Islamic Rep. of Iran 1.8 7.3
Canada 1.8 15.4
Korea 1.8 10.6
United Kingdom 1.6 7.5

Cumulative

Top-10 cumulative energy-related CO2 emitters between 1850-2008[82]
Country  % of world
total
Metric tonnes
CO2 per person
United States 28.5 1,132.7
China 9.36 85.4
Russian Federation 7.95 677.2
Germany 6.78 998.9
United Kingdom 5.73 1,127.8
Japan 3.88 367
France 2.73 514.9
India 2.52 26.7
Canada 2.17 789.2
Ukraine 2.13 556.4

Embedded emissions

One way of attributing greenhouse gas (GHG) emissions is to measure the embedded emissions (also referred to as "embodied emissions") of goods that are being consumed. Emissions are usually measured according to production, rather than consumption.[83] For example, in the main international treaty on climate change (the UNFCCC), countries report on emissions produced within their borders, e.g., the emissions produced from burning fossil fuels.[75]:179[84]:1 Under a production-based accounting of emissions, embedded emissions on imported goods are attributed to the exporting, rather than the importing, country. Under a consumption-based accounting of emissions, embedded emissions on imported goods are attributed to the importing country, rather than the exporting, country.
Davis and Caldeira (2010)[84]:4 found that a substantial proportion of CO2 emissions are traded internationally. The net effect of trade was to export emissions from China and other emerging markets to consumers in the US, Japan, and Western Europe. Based on annual emissions data from the year 2004, and on a per-capita consumption basis, the top-5 emitting countries were found to be (in tCO2 per person, per year): Luxembourg (34.7), the US (22.0), Singapore (20.2), Australia (16.7), and Canada (16.6).[84]:5

Effect of policy

Rogner et al. (2007) assessed the effectiveness of policies to reduce emissions (mitigation of climate change).[61] They concluded that mitigation policies undertaken by UNFCCC Parties were inadequate to reverse the trend of increasing GHG emissions. The impacts of population growth, economic development, technological investment, and consumption had overwhelmed improvements in energy intensities and efforts to decarbonize (energy intensity is a country's total primary energy supply (TPES) per unit of GDP (Rogner et al., 2007).[85] TPES is a measure of commercial energy consumption (World Bank, 2010, p. 371)).[64]

Projections

Based on then-current energy policies, Rogner et al. (2007) projected that energy-related CO2 emissions in 2030 would be 40-110% higher than in 2000.[61] Two-thirds of this increase was projected to come from non-Annex I countries. Per capita emissions in Annex I countries were still projected to remain substantially higher than per capita emissions in non-Annex I countries. Projections consistently showed a 25-90% increase in the Kyoto gases (carbon dioxide, methane, nitrous oxide, sulphur hexafluoride) compared to 2000.
IEA (2007, p. 199) estimated future cumulative energy-related CO2 emissions for several countries.[67] Their reference scenario projected cumulative energy-related CO2 emissions between the years 1900 and 2030. In this scenario, China’s share of cumulative emissions rises to 16%, approaching that of the United States (25%) and the European Union (18%). India’s cumulative emissions (4%) approach those of Japan (4%).

Relative CO2 emission from various fuels

One liter of gasoline, when used as a fuel, produces 2.32 kg (1.3 cubic meters) of carbon dioxide, a greenhouse gas. One US gallon produces 19.4 lb (172.65 cubic feet)[86][87][88]
Mass of carbon dioxide emitted per quantity of energy for various fuels[89]
Fuel name CO2
emitted
(lbs/106 Btu)
CO2
emitted
(g/106 J)
Natural gas 117 50.30
Liquefied petroleum gas 139 59.76
Propane 139 59.76
Aviation gasoline 153 65.78
Automobile gasoline 156 67.07
Kerosene 159 68.36
Fuel oil 161 69.22
Tires/tire derived fuel 189 81.26
Wood and wood waste 195 83.83
Coal (bituminous) 205 88.13
Coal (sub-bituminous) 213 91.57
Coal (lignite) 215 92.43
Petroleum coke 225 96.73
Coal (anthracite) 227 97.59