You are on page 1of 12

Greenhouse gas

From Wikipedia, the free encyclopedia


Jump to navigationJump to search

The greenhouse effect of solar radiation on the Earth's surface caused by


greenhouse gases
A greenhouse gas (sometimes abbreviated GHG) is a gas that absorbs and emits
radiant energy within the thermal infrared range. Greenhouse gases cause the
greenhouse effect[1] on planets. The primary greenhouse gases in Earth's atmosphere
are water vapor (H
2O), carbon dioxide (CO
2), methane (CH
4), nitrous oxide (N
2O), and ozone (O3). Without greenhouse gases, the average temperature of Earth's
surface would be about -18 �C (0 �F),[2] rather than the present average of 15 �C
(59 �F).[3][4][5] The atmospheres of Venus, Mars and Titan also contain greenhouse
gases.

Human activities since the beginning of the Industrial Revolution (around 1750)
have produced a 45% increase in the atmospheric concentration of carbon dioxide,
from 280 ppm in 1750 to 415 ppm in 2019.[6] The last time the atmospheric
concentration of carbon dioxide was this high was over 3 million years ago. [7]
This increase has occurred despite the uptake of more than half of the emissions by
various natural "sinks" involved in the carbon cycle.[8][9] The vast majority of
anthropogenic carbon dioxide emissions come from combustion of fossil fuels,
principally coal, oil, and natural gas, with additional contributions coming from
deforestation, changes in land use, soil erosion and agriculture (including
livestock).[10][11] The leading source of anthropogenic methane emissions is animal
agriculture, followed by fugitive emissions from gas, oil, coal and other industry,
solid waste, wastewater and rice production.[12]

At current emission rates, temperatures could increase by 2 �C (3.6�F), which the


United Nations' Intergovernmental Panel on Climate Change (IPCC) designated as the
upper limit to avoid "dangerous" levels, by 2036.[13]

Contents
1 Gases in Earth's atmosphere
1.1 Non-greenhouse gases
1.2 Greenhouse gases
1.3 Indirect radiative effects
1.4 Contribution of clouds to Earth's greenhouse effect
2 Impacts on the overall greenhouse effect
2.1 Proportion of direct effects at a given moment
2.2 Atmospheric lifetime
2.3 Radiative forcing
2.4 Global warming potential
3 Natural and anthropogenic sources
3.1 Ice cores
3.2 Changes since the Industrial Revolution
4 Role of water vapor
5 Anthropogenic greenhouse gases
5.1 Greenhouse gases emissions by sector
5.2 Regional and national attribution of emissions
5.3 From land-use change
5.4 Greenhouse gas intensity
5.5 Cumulative and historical emissions
5.6 Changes since a particular base year
5.7 Annual emissions
5.8 Top emitter countries
5.9 Embedded emissions
5.10 Effect of policy
5.11 Projections
5.12 Relative CO2 emission from various fuels
6 Life-cycle greenhouse-gas emissions of energy sources
7 Removal from the atmosphere
7.1 Natural processes
7.2 Negative emissions
8 History of scientific research
9 See also
10 References
11 Bibliography
12 External links
12.1 Carbon dioxide emissions
Gases in Earth's atmosphere
Main articles: Greenhouse effect and Atmosphere of Earth
Non-greenhouse gases
The major constituents of Earth's atmosphere, nitrogen (N
2)(78%), oxygen (O
2)(21%), and argon (Ar)(0.9%), are not greenhouse gases because molecules
containing two atoms of the same element such as N
2 and O
2 have no net change in the distribution of their electrical charges when they
vibrate, and monatomic gases such as Ar do not have vibrational modes. Hence they
are almost totally unaffected by infrared radiation. Some molecules containing just
two atoms of different elements, such as carbon monoxide (CO) and hydrogen chloride
(HCl), do absorb infrared radiation, but these molecules are short-lived in the
atmosphere owing to their reactivity or solubility. Therefore, they do not
contribute significantly to the greenhouse effect and often are omitted when
discussing greenhouse gases.

Greenhouse gases
refer to caption and adjacent text
Atmospheric absorption and scattering at different wavelengths of electromagnetic
waves. The largest absorption band of carbon dioxide is not far from the maximum in
the thermal emission from ground, and it partly closes the window of transparency
of water; hence its major effect.
Greenhouse gases are those that absorb and emit infrared radiation in the
wavelength range emitted by Earth.[1] Carbon dioxide (0.04%), nitrous oxide,
methane and ozone are trace gases that account for almost one tenth of 1% of
Earth's atmosphere and have an appreciable greenhouse effect.

See also: Global warming and Carbon dioxide in Earth's atmosphere


In order, the most abundant[clarification needed] greenhouse gases in Earth's
atmosphere are:[citation needed]

Water vapor (H
2O)
Carbon dioxide (CO
2)
Methane (CH
4)
Nitrous oxide (N
2O)
Ozone (O
3)
Chlorofluorocarbons (CFCs)
Hydrofluorocarbons (includes HCFCs and HFCs)
Atmospheric concentrations 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 or
absorption by bodies of water).[14] The proportion of an emission remaining in the
atmosphere after a specified time is the "airborne fraction" (AF). The annual
airborne fraction is the ratio of the atmospheric increase in a given year to that
year's total emissions. As of 2006 the annual airborne fraction for CO
2 was about 0.45. The annual airborne fraction increased at a rate of 0.25 � 0.21%
per year over the period 1959�2006.[15]

Indirect radiative effects


world map of carbon monoxide concentrations in the lower atmosphere
The false colors in this image represent concentrations of carbon monoxide in the
lower atmosphere, ranging from about 390 parts per billion (dark brown pixels), to
220 parts per billion (red pixels), to 50 parts per billion (blue pixels).[16]
Some gases have indirect radiative effects (whether or not they are greenhouse
gases themselves). This happens in two main ways. One way is that when they break
down in the atmosphere they produce another greenhouse gas. For example, methane
and carbon monoxide (CO) are oxidized to give carbon dioxide (and methane oxidation
also produces water vapor). Oxidation of CO to CO
2 directly produces an unambiguous increase in radiative forcing although the
reason is subtle. The peak of the thermal IR emission from Earth's surface is very
close to a strong vibrational absorption band of CO
2 (wavelength 15 microns, or wavenumber 667 cm-1). On the other hand, the single CO
vibrational band only absorbs IR at much shorter wavelengths (4.7 microns, or 2145
cm-1), where the emission of radiant energy from Earth's surface is at least a
factor of ten lower. Oxidation of methane to CO
2, which requires reactions with the OH radical, produces an instantaneous
reduction in radiative absorption and emission since CO
2 is a weaker greenhouse gas than methane. However, the oxidations of CO and CH
4 are entwined since both consume OH radicals. In any case, the calculation of the
total radiative effect includes both direct and indirect forcing.

A second type of indirect effect happens when chemical reactions in the atmosphere
involving these gases change the concentrations of greenhouse gases. For example,
the destruction of non-methane volatile organic compounds (NMVOCs) in the
atmosphere can produce ozone. The size of the indirect effect can depend strongly
on where and when the gas is emitted.[17]

Methane has indirect effects in addition to forming CO


2. The main chemical that reacts with methane in the atmosphere is the hydroxyl
radical (OH), thus more methane means that the concentration of OH goes down.
Effectively, methane increases its own atmospheric lifetime and therefore its
overall radiative effect. The oxidation of methane can produce both ozone and
water; and is a major source of water vapor in the normally dry stratosphere. CO
and NMVOCs produce CO
2 when they are oxidized. They remove OH from the atmosphere, and this leads to
higher concentrations of methane. The surprising effect of this is that the global
warming potential of CO is three times that of CO
2.[18] The same process that converts NMVOCs to carbon dioxide can also lead to the
formation of tropospheric ozone. Halocarbons have an indirect effect because they
destroy stratospheric ozone. Finally, hydrogen can lead to ozone production and CH
4 increases as well as producing stratospheric water vapor.[17]

Contribution of clouds to Earth's greenhouse effect


The major non-gas contributor to Earth's greenhouse effect, clouds, also absorb and
emit infrared radiation and thus have an effect on greenhouse gas radiative
properties. Clouds are water droplets or ice crystals suspended in the atmosphere.
[19][20]
Impacts on the overall greenhouse effect
refer to caption and adjacent text
Schmidt et al. (2010)[21] analysed how individual components of the atmosphere
contribute to the total greenhouse effect. They estimated that water vapor accounts
for about 50% of Earth's greenhouse effect, with clouds contributing 25%, carbon
dioxide 20%, and the minor greenhouse gases and aerosols accounting for the
remaining 5%. In the study, the reference model atmosphere is for 1980 conditions.
Image credit: NASA.[22]
The contribution of each gas to the greenhouse effect is determined by the
characteristics of that gas, its abundance, and any indirect effects it may cause.
For example, the direct radiative effect of a mass of methane is about 84 times
stronger than the same mass of carbon dioxide over a 20-year time frame[23] but it
is present in much smaller concentrations so that its total direct radiative effect
has so far been smaller, in part due to its shorter atmospheric lifetime in the
absence of additional carbon sequestration. 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)[24] argue that the
contribution to climate change from methane is at least double previous estimates
as a result of this effect.[25]

When ranked by their direct contribution to the greenhouse effect, the most
important are:[19][failed verification]

Compound
Formula
Concentration in
atmosphere[26] (ppm) Contribution
(%)
Water vapor and clouds H
2O 10�50,000(A) 36�72%
Carbon dioxide CO
2 ~400 9�26%
Methane CH
4 ~1.8 4�9%
Ozone O
3 2�8(B) 3�7%
notes:
(A) Water vapor strongly varies locally[27]
(B) The concentration in stratosphere. About 90% of the ozone in Earth's atmosphere
is contained in the stratosphere.

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.[28]

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.[19][20]
In addition, some gases, such as methane, are known to have large indirect effects
that are still being quantified.[29]

Atmospheric lifetime
Aside from water vapor, which has a residence time of about nine days,[30] major
greenhouse gases are well mixed and take many years to leave the atmosphere.[31]
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)[32] defines the lifetime {\displaystyle \tau }\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 {\displaystyle \tau }\tau can be defined as the ratio of
the mass {\displaystyle m}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 ({\displaystyle F_{out}}F_{{out}}),
chemical loss of X ({\displaystyle L}L), and deposition of X ({\displaystyle D}D)
(all in kg/s): {\displaystyle \tau ={\frac {m}{F_{out}+L+D}}}\tau ={\frac {m}
{F_{{out}}+L+D}}.[32] If output of this gas into the box ceased, then after time
{\displaystyle \tau }\tau , its concentration would decrease by about 63%.

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.[33] The atmospheric lifetime of CO
2 is estimated of the order of 30�95 years.[34] This figure accounts for CO
2 molecules being removed from the atmosphere by mixing into the ocean,
photosynthesis, and other processes. However, this excludes the balancing fluxes of
CO
2 into the atmosphere from the geological reservoirs, which have slower
characteristic rates.[35] Although more than half of the CO
2 emitted is removed from the atmosphere within a century, some fraction (about
20%) of emitted CO
2 remains in the atmosphere for many thousands of years.[36] [37] [38] Similar
issues apply to other greenhouse gases, many of which have longer mean lifetimes
than CO
2, e.g. N2O has a mean atmospheric lifetime of 121 years.[23]

Radiative forcing
Earth absorbs some of the radiant energy received from the sun, reflects some of it
as light and reflects or radiates the rest back to space as heat.[39] Earth's
surface temperature depends on this balance between incoming and outgoing energy.
[39] If this energy balance is shifted, Earth's surface becomes warmer or cooler,
leading to a variety of changes in global climate.[39]

A number of natural and man-made mechanisms can affect the global energy balance
and force changes in Earth's climate.[39] Greenhouse gases are one such mechanism.
[39] Greenhouse gases absorb and emit some of the outgoing energy radiated from
Earth's surface, causing that heat to be retained in the lower atmosphere.[39] As
explained above, some greenhouse gases remain in the atmosphere for decades or even
centuries, and therefore can affect Earth's energy balance over a long period.[39]
Radiative forcing quantifies the effect of factors that influence Earth's energy
balance, including changes in the concentrations of greenhouse gases.[39] Positive
radiative forcing leads to warming by increasing the net incoming energy, whereas
negative radiative forcing leads to cooling.[39]

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 CO
2 and evaluated for a specific timescale. Thus, if a gas has a high (positive)
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 CO
2 its GWP will increase when the timescale is considered. Carbon dioxide is defined
to have a GWP of 1 over all time periods.

Methane has an atmospheric lifetime of 12 � 3 years. The 2007 IPCC report lists the
GWP as 72 over a time scale of 20 years, 25 over 100 years and 7.6 over 500 years.
[40] A 2014 analysis, however, states that although methane's initial impact is
about 100 times greater than that of CO
2, because of the shorter atmospheric lifetime, after six or seven decades, the
impact of the two gases is about equal, and from then on methane's relative role
continues to decline.[41] The decrease in GWP at longer times is because methane is
degraded to water and CO
2 through chemical reactions in the atmosphere.

Examples of the atmospheric lifetime and GWP relative to CO


2 for several greenhouse gases are given in the following table:

Atmospheric lifetime and GWP relative to CO


2 at different time horizon for various greenhouse gases
Gas name Chemical
formula Lifetime
(years)[23] Global warming potential (GWP) for given time horizon
20-yr[23] 100-yr[23] 500-yr[40]
Carbon dioxide CO
2 30�95 1 1 1
Methane CH
4 12 84 28 7.6
Nitrous oxide N
2O 121 264 265 153
CFC-12 CCl
2F
2 100 10 800 10 200 5 200
HCFC-22 CHClF
2 12 5 280 1 760 549
Tetrafluoromethane CF
4 50 000 4 880 6 630 11 200
Hexafluoroethane C
2F
6 10 000 8 210 11 100 18 200
Sulfur hexafluoride SF
6 3 200 17 500 23 500 32 600
Nitrogen trifluoride NF
3 500 12 800 16 100 20 700
The use of CFC-12 (except some essential uses) has been phased out due to its ozone
depleting properties.[42] The phasing-out of less active HCFC-compounds will be
completed in 2030.[43]

Carbon dioxide in Earth's atmosphere if half of global-warming emissions[44][45]


are not absorbed.
(NASA simulation; 9 November 2015)
Natural and anthropogenic sources
refer to caption and article text
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.
refer to caption and image description
This diagram shows a simplified representation of the contemporary global carbon
cycle. Changes are measured in gigatons of carbon per year (GtC/y). Canadell et al.
(2007) estimated the growth rate of global average atmospheric CO
2 for 2000�2006 as 1.93 parts-per-million per year (4.1 petagrams of carbon per
year).[46]
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, because the large natural
sources and sinks roughly balanced. In the industrial era, human activities have
added greenhouse gases to the atmosphere, mainly through the burning of fossil
fuels and clearing of forests.[47][48]

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".[49] In AR4, "most of" is defined as more than 50%.

Abbreviations used in the two tables below: ppm = parts-per-million; ppb = parts-
per-billion; ppt = parts-per-trillion; W/m2 = watts per square metre

Current greenhouse gas concentrations[50]


Gas Pre-1750
tropospheric
concentration[51] Recent
tropospheric
concentration[52] Absolute increase
since 1750 Percentage
increase
since 1750 Increased
radiative forcing
(W/m2)[53]
Carbon dioxide (CO
2) 280 ppm[54] 395.4 ppm[55] 115.4 ppm 41.2% 1.88
Methane (CH
4) 700 ppb[56] 1893 ppb /[57][58]
1762 ppb[57] 1193 ppb /
1062 ppb 170.4% /
151.7% 0.49
Nitrous oxide (N
2O) 270 ppb[53][59] 326 ppb /[57]
324 ppb[57] 56 ppb /
54 ppb 20.7% /
20.0% 0.17
Tropospheric
ozone (O
3) 237 ppb[51] 337 ppb[51] 100 ppb 42% 0.4[60]
Relevant to radiative forcing and/or ozone depletion; all of the following have no
natural sources and hence zero amounts pre-industrial[50]
Gas Recent
tropospheric
concentration Increased
radiative forcing
(W/m2)
CFC-11
(trichlorofluoromethane)
(CCl
3F) 236 ppt /
234 ppt 0.061
CFC-12 (CCl
2F
2) 527 ppt /
527 ppt 0.169
CFC-113 (Cl
2FC-CClF
2) 74 ppt /
74 ppt 0.022
HCFC-22 (CHClF
2) 231 ppt /
210 ppt 0.046
HCFC-141b (CH
3CCl
2F) 24 ppt /
21 ppt 0.0036
HCFC-142b (CH
3CClF
2) 23 ppt /
21 ppt 0.0042
Halon 1211 (CBrClF
2) 4.1 ppt /
4.0 ppt 0.0012
Halon 1301 (CBrClF
3) 3.3 ppt /
3.3 ppt 0.001
HFC-134a (CH
2FCF
3) 75 ppt /
64 ppt 0.0108
Carbon tetrachloride (CCl
4) 85 ppt /
83 ppt 0.0143
Sulfur hexafluoride (SF
6) 7.79 ppt /[61]
7.39 ppt[61] 0.0043
Other halocarbons Varies by
substance collectively
0.02
Halocarbons in total 0.3574
refer to caption and article text
400,000 years of ice core data
Ice cores provide evidence for greenhouse gas concentration variations over the
past 800,000 years (see the following section). Both CO
2 and CH
4 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 that indicates CO
2 mole fractions stayed within a range of 180 ppm to 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 CO
2 levels were likely 10 times higher than now.[62] Indeed, higher CO
2 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.[63][64][65]
The spread of land plants is thought to have reduced CO
2 concentrations during the late Devonian, and plant activities as both sources and
sinks of CO
2 have since been important in providing stabilising feedbacks.[66] 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 that raised the CO
2 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.[67] 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 approximately 0.645 billion tonnes
of CO
2 per year, whereas humans contribute 29 billion tonnes of CO
2 each year.[68][67][69][70]

Ice cores
Measurements from Antarctic ice cores show that before industrial emissions started
atmospheric CO
2 mole fractions were about 280 parts per million (ppm), and stayed between 260 and
280 during the preceding ten thousand years.[71] 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,[72] though others have argued that these findings more likely reflect
calibration or contamination problems rather than actual CO
2 variability.[73][74] 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.

Changes since the Industrial Revolution


Refer to caption
Recent year-to-year increase of atmospheric CO
2.
Refer to caption
Major greenhouse gas trends.
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 to 415 ppm, or 120 ppm over modern pre-industrial
levels. The first 30 ppm increase took place in about 200 years, from the start of
the Industrial Revolution to 1958; however the next 90 ppm increase took place
within 56 years, from 1958 to 2014.[75][76]

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.[77]

Total cumulative emissions from 1870 to 2017 were 425�20 GtC (1539 GtCO2) from
fossil fuels and industry, and 180�60 GtC (660 GtCO2) from land use change. Land-
use change, such as deforestation, caused about 31% of cumulative emissions over
1870�2017, coal 32%, oil 25%, and gas 10%.[78]

Today,[when?] 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.[79]

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.

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.
[20] Water vapor concentrations fluctuate regionally, but human activity does not
directly affect water vapor concentrations except at local scales, such as near
irrigated fields. Indirectly, human activity that increases global temperatures
will increase water vapor concentrations, a process known as water vapor feedback.
[80] The atmospheric concentration of vapor is highly variable and depends largely
on temperature, from less than 0.01% in extremely cold regions up to 3% by mass in
saturated air at about 32 �C.[81] (See Relative humidity#other important facts.)

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 CH
4 and CO
2.[82] Water vapor responds to and amplifies effects of the other greenhouse gases.
The Clausius�Clapeyron relation establishes that more water vapor will be present
per unit volume at elevated temperatures. 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 (assuming that the relative
humidity remains approximately constant; modeling and observational studies find
that this is indeed so). 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.[80]

Anthropogenic greenhouse gases


See also: Climate change and ecosystems

This graph shows changes in the annual greenhouse gas index (AGGI) between 1979 and
2011. [83] The AGGI measures the levels of greenhouse gases in the atmosphere based
on their ability to cause changes in Earth's climate.[83]

This bar graph shows global greenhouse gas emissions by sector from 1990 to 2005,
measured in 100-year estimated carbon dioxide equivalents.[84]

Modern global CO2 emissions from the burning of fossil fuels.


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.[85] Natural sources of
carbon dioxide are more than 20 times greater than sources due to human activity,
[86] 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.[87]

It is likely that anthropogenic (i.e., human-induced) warming, such as that due to


elevated greenhouse gas levels, has had a discernible influence on many physical
and biological systems.[88] Future warming is projected to have a range of impacts,
including sea level rise,[89] increased frequencies and severities of some extreme
weather events,[89] loss of biodiversity,[90] and regional changes in agricultural
productivity.[90]
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 CO
2 emissions.[87]
livestock enteric fermentation and manure management,[91] paddy rice farming, land
use and wetland changes, man-made lakes,[92] 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 (N
2O) concentrations.
Mean greenhouse gas emissions for different food types[93]
Food Types Greenhouse Gas Emissions (g CO2-Ceq per g protein)
Ruminant Meat
62
Recirculating Aquaculture
30
Trawling Fishery
26
Non-recirculating Aquaculture
12
Pork
10
Poultry
10
Dairy
9.1
Non-trawling Fishery
8.6
Eggs
6.8
Starchy Roots
1.7
Wheat
1.2
Maize
1.2
Legumes
0.25
The seven sources of CO
2 from fossil fuel combustion are (with percentage contributions for 2000�2004):
[94]

This list needs updating, as it uses an out of date source.[needs update]

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 %
Carbon dioxide, methane, nitrous oxide (N
2O) and three groups of fluorinated gases (sulfur hexafluoride (SF
6), hydrofluorocarbons (HFCs), and perfluorocarbons (PFCs)) are the major
anthropogenic greenhouse gases,[95]:147[96] and are regulated under the Kyoto
Protocol international treaty, which came into force in 2005.[97] Emissions
limitations specified in the Kyoto Protocol expired in 2012.[97] The Canc�n
agreement, agreed on in 2010, includes voluntary pledges made by 76 countries to
control emissions.[98] At the time of the agreement, these 76 countries were
collectively responsible for 85% of annual global emissions.[98]

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 15
October 2016, negotiators from over 170 nations meeting at the summit of the United
Nations Environment Programme reached a legally binding accord to phase out
hydrofluorocarbons (HFCs) in an amendment to the Montreal Protocol.[99][100][101]

Chart showing 2016 global greenhouse gas emissions by sector[102]. Percentages are
calculated from estimated global emissions of all Kyoto Greenhouse Gases, converted
to CO2 equivalent quantities (GtCO2e).
Greenhouse gases emissions by sector
Global greenhouse gas emissions can be attributed to different sectors of the
economy. This provides a picture of the varying contributions of different types of
economic activity to global warming, and helps in understanding the changes
required to mitigate climate change.

Manmade greenhouse gas emissions can be divided into those that arise from the
combustion of fuels to produce energy, and those generated by other processes.
Around two thirds of greenhouse gas emissions arise from the combustion of
fuels[103].

Energy may be produced at the point of consumption, or by a generator for


consumption by others. Thus emissions arising from energy production may be
categorised according to where they are emitted, or where the resulting energy is
consumed. If emissions are attributed at the point of production, then electricity
generators contribute about 25% of global greenhouse gas emissions[104]. If these
emissions are attributed to the final consumer then 24% of total emissions arise
from manufacturing and construction, 17% from transportation, 11% from domestic
consumers, and 7% from commercial consumers[105]. Around 4% of emissions arise from
the energy consumed by the energy and fuel industry itself.

The remaining third of emissions arise from processes other than energy production.
12% of total emissions arise from agriculture, 7% from land use change and
forestry, 6% from industrial processes, and 3% from waste[103] . Around 6% of
emissions are fugitive emissions, which are waste gases released by the extraction
of

You might also like