To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Methane emissions

From Wikipedia, the free encyclopedia

Sources of methane emissions due to human activity (year 2020 estimates) [1]

  Fossil Fuel Use (33%)
  Animal Agriculture (30%)
  Plant Agriculture (18%)
  Waste (15%)
  All Other (4%)

Increasing methane emissions are a major contributor to the rising concentration of greenhouse gases in Earth's atmosphere, and are responsible for up to one-third of near-term global heating.[1][2] During 2019, about 60% (360 million tons) of methane released globally was from human activities, while natural sources contributed about 40% (230 million tons).[3][4] Reducing methane emissions by capturing and utilizing the gas can produce simultaneous environmental and economic benefits.[1][5]

Since the Industrial Revolution, concentrations of methane in the atmosphere have more than doubled, and about 20 percent of the warming the planet has experienced can be attributed to the gas.[6] About one-third (33%) of anthropogenic emissions are from gas release during the extraction and delivery of fossil fuels; mostly due to gas venting and gas leaks from both active fossil fuel infrastructure and orphan wells.[7] Russia is the world's top methane emitter from oil and gas.[8][9]

Animal agriculture is a similarly large source (30%); primarily because of enteric fermentation by ruminant livestock such as cattle and sheep. According to the Global Methane Assessment published in 2021, methane emissions from livestock (including cattle) are the largest sources of agricultural emissions worldwide[10] A single cow can make up to 99 kg of methane gas per year.[11] Ruminant livestock can produce 250 to 500 L of methane per day.[12]

Human consumer waste flows, especially those passing through landfills and wastewater treatment, have grown to become a third major category (18%). Plant agriculture, including both food and biomass production, constitutes a fourth group (15%), with rice production being the largest single contributor.[1][13]

The world's wetlands contribute about three-quarters (75%) of the enduring natural sources of methane.[3][4] Seepages from near-surface hydrocarbon and clathrate hydrate deposits, volcanic releases, wildfires, and termite emissions account for much of the remainder.[13] Contributions from the surviving wild populations of ruminant mammals are vastly overwhelmed by those of cattle, humans, and other livestock animals.[14]

The Economist recommended setting methane emissions targets as a reduction in methane emissions would allow for more time to tackle the more challenging carbon emissions".[15][16]

YouTube Encyclopedic

  • 1/5
    Views:
    134 604
    8 830
    19 324
    2 168
    5 206
  • This climate gas is scarier than CO2
  • Large landfills in India & South Asia generating big methane Clouds. What to do? UPSC
  • Methane Emergency
  • Remote sensing of methane emissions: Results and possibilities
  • How are agriculture and livestock responsible for GHG emissions? | InDepth | Drishti IAS English

Transcription

Atmospheric concentration and warming influence

Globally averaged atmospheric concentration and its annual growth rate.[17] In April 2022, NOAA reported an annual increase in global atmospheric methane of 17 parts per billion (ppb) in 2021—averaging 1,895.7 ppb in that year—the largest annual increase recorded since systematic measurements began in 1983; the increase during 2020 was 15.3 ppb, itself a record increase.[18]

The atmospheric methane (CH4) concentration is increasing and exceeded 1860 parts per billion in 2019, equal to two-and-a-half times the pre-industrial level.[19] The methane itself causes direct radiative forcing that is second only to that of carbon dioxide (CO2).[20] Due to interactions with oxygen compounds stimulated by sunlight, CH4 can also increase the atmospheric presence of shorter-lived ozone and water vapour, themselves potent warming gases: atmospheric researchers call this amplification of methane's near-term warming influence indirect radiative forcing.[21] When such interactions occur, longer-lived and less-potent CO2 is also produced. Including both the direct and indirect forcings, the increase in atmospheric methane is responsible for about one-third of near-term global heating.[1][2]

Though methane causes far more heat to be trapped than the same mass of carbon dioxide, less than half of the emitted CH4 remains in the atmosphere after a decade. On average, carbon dioxide warms for much longer, assuming no change in rates of carbon sequestration.[22][23] The global warming potential (GWP) is a way of comparing the warming due to other gases to that from carbon dioxide, over a given time period. Methane's GWP20 of 85 means that a ton of CH4 emitted into the atmosphere creates approximately 85 times the atmospheric warming as a ton of CO2 over a period of 20 years.[23] On a 100-year timescale, methane's GWP100 is in the range of 28–34.

Methane emissions are important as reducing them can buy time to tackle carbon emissions.[24][25]

Overview of emission sources

The main sources of methane for the decade 2008–2017, estimated by the Global Carbon Project[17]
"Methane global emissions from the five broad categories for the 2008–2017 decade for top-down inversion models and for bottom-up models and inventories (right dark coloured box plots).[17][clarification needed]

Biogenic methane is actively produced by microorganisms in a process called methanogenesis. Under certain conditions, the process mix responsible for a sample of methane may be deduced from the ratio of the isotopes of carbon, and through analysis methods similar to carbon dating.[26][27]

Anthropogenic

Map of methane emissions from four source categories[17]

As of 2020, emission volumes from some sources remain more uncertain than others; due in part to localized emission spikes not captured by the limited global measurement capability. The time required for a methane emission to become well-mixed throughout earth's troposphere is about 1–2 years.[28]

Satellite data indicate over 80% of the growth of methane emissions during 2010–2019 are tropical terrestrial emissions.[29][30]

There is accumulating research and data showing that oil and gas industry methane emissions – or from fossil fuel extraction, distribution and use – are much larger than thought.[31][32][33][34][35]

Category Major Sources IEA Annual Emission[3]
(Million Tons)
Fossil fuels Gas distribution 45
Oil wells 39*
Coal mines 39
Biofuels Anaerobic digestion 11
Industrial agriculture Enteric fermentation 145
Rice paddies
Manure management
Biomass Biomass burning 16
Consumer waste Solid waste
Landfill gas
68
Wastewater
Total anthropogenic 363
* An additional 100 million tons (140 billion cubic meters) of gas is flared each year from oil wells.[36]
Additional References: [1][37][38][39][40]

Natural

Map of methane emissions from three natural sources and one sink.[17]

Natural sources have always been a part of the methane cycle. Wetland emissions have been declining due to draining for agricultural and building areas.

Category Major Sources IEA Annual Emission[3]
(Million Tons)
Wetlands Wetland methane 194
Other natural Geologic seepages
Volcanic gas
39
Arctic methane emissions
Ocean sediments
Wildfires
Termites
Total natural 233
Additional References: [1][37][38]

Methanogenesis

Most ecological emissions of methane relate directly to methanogens generating methane in warm, moist soils as well as in the digestive tracts of certain animals. Methanogens are methane producing microorganisms. In order to produce energy, they use an anaerobic process called methanogenesis. This process is used in lieu of aerobic, or with oxygen, processes because methanogens are unable to metabolise in the presence of even small concentrations of oxygen. When acetate is broken down in methanogenesis, the result is the release of methane into the surrounding environment.

Methanogenesis, the scientific term for methane production, occurs primarily in anaerobic conditions because of the lack of availability of other oxidants. In these conditions, microscopic organisms called archaea use acetate and hydrogen to break down essential resources[vague] in a process called fermentation.

Acetoclastic methanogenesis – certain archaea cleave acetate produced during anaerobic fermentation to yield methane and carbon dioxide.

H3C-COOH → CH4 + CO2

Hydrogenotrophic methanogenesis – archaea oxidize hydrogen with carbon dioxide to yield methane and water.

4H2 + CO2 → CH4 + 2H2O

While acetoclastic methanogenesis and hydrogenotrophic methanogenesis are the two major source reactions for atmospheric methane, other minor biological methane source reactions also occur. For example, it has been discovered that leaf surface wax exposed to UV radiation in the presence of oxygen is an aerobic source of methane.[41]

Natural methane cycles

Methane observations from 2005 to 2014 showing the seasonal variations and the difference between northern and southern hemispheres

Emissions of methane into the atmosphere are directly related to temperature and moisture. Thus, the natural environmental changes that occur during seasonal change act as a major control of methane emission. Additionally, even changes in temperature during the day can affect the amount of methane that is produced and consumed.[citation needed]

Its concentration is higher in the Northern Hemisphere since most sources (both natural and human) are located on land and the Northern Hemisphere has more land mass.[42] The concentrations vary seasonally, with, for example, a minimum in the northern tropics during April−May mainly due to removal by the hydroxyl radical.[43]

For example, plants that produce methane can emit as much as two to four times more methane during the day than during the night.[44] This is directly related to the fact that plants tend to rely on solar energy to enact chemical processes.

Additionally, methane emissions are affected by the level of water sources. Seasonal flooding during the spring and summer naturally increases the amount of methane released into the air.[citation needed]

Wetlands

Greenhouse gas emissions from wetlands of concern consist primarily of methane and nitrous oxide emissions. Wetlands are the largest natural source of atmospheric methane in the world, and are therefore a major area of concern with respect to climate change.[45][46][47] Wetlands account for approximately 20 - 30% of atmospheric methane through emissions from soils and plants, and contribute an approximate average of 161 Tg of methane to the atmosphere per year.[48]

Wetlands are characterized by water-logged soils and distinctive communities of plant and animal species that have adapted to the constant presence of water. This high level of water saturation creates conditions conducive to methane production. Most methanogenesis, or methane production, occurs in oxygen-poor environments. Because the microbes that live in warm, moist environments consume oxygen more rapidly than it can diffuse in from the atmosphere, wetlands are the ideal anaerobic environments for fermentation as well as methanogen activity. However, levels of methanogenesis fluctuates due to the availability of oxygen, soil temperature, and the composition of the soil. A warmer, more anaerobic environment with soil rich in organic matter would allow for more efficient methanogenesis.[49]

In wetlands, where the rate of methane production is high, plants help methane travel into the atmosphere—acting like inverted lightning rods as they direct the gas up through the soil and into the air. They are also suspected to produce methane themselves, but because the plants would have to use aerobic conditions to produce methane, the process itself is still unidentified, according to a 2014 Biogeochemistry article.[50]

A 1994 article on methane emissions from northern wetlands said that since the 1800s, atmospheric methane concentrations increased annually at a rate of about 0.9%.[44]

Human-caused methane emissions

The AR6 of the IPCC said, "It is unequivocal that the increases in atmospheric carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O) since the pre-industrial period are overwhelmingly caused by human activities."[51][52][53] Atmospheric methane accounted for 20% of the total radiative forcing (RF) from all of the long-lived and globally mixed greenhouse gases.

According to the 2021 assessment by the Climate and Clean Air Coalition (CCAC) and the United Nations Environment Programme (UNEP) over 50% of global methane emissions are caused by human activities in fossil fuels (35%), waste (20%), and agriculture (40%). The oil and gas industry accounts for 23%, and coal mining for 12%. Twenty percent of global anthropogenic emissions stem from landfills and wastewater. Manure and enteric fermentation represent 32%, and rice cultivation represents 8%.[54]

The most clearly identified rise in atmospheric methane as a result of human activity occurred in the 1700s during the industrial revolution. During the 20th century—mainly because of the use of fossil fuels—concentration of methane in the atmosphere increased, then stabilized briefly in the 1990s,[55] only to begin to increase again in 2007. After 2014, the increase accelerated and by 2017, reached 1,850 (parts per billion) ppb.[56][57]

Increases in methane levels due to modern human activities arise from a number of specific sources including industrial activity; from extraction of oil and natural gas from underground reserves;[58] transportation via pipeline of oil and natural gas; and melting permafrost in Arctic regions, due to global warming which is caused by human use of fossil fuels.

The primary component of natural gas is methane, which is emitted to the atmosphere in every stage of natural gas "production, processing, storage, transmission, and distribution".[59]

Emissions due to oil and gas extraction

A 2005 Wuppertal Institute for Climate, Environment and Energy article identified pipelines that transport natural gas as a source of methane emissions. The article cited the example of Trans-Siberian natural gas pipeline system to western and Central Europe from the Yamburg and Urengoy exist gas fields in Russia with a methane concentration of 97%.[60] In accordance with the IPCC and other natural gas emissions control groups, measurements had to be taken throughout the pipeline to measure methane emissions from technological discharges and leaks at the pipeline fittings and vents. Although the majority of the natural gas leaks were carbon dioxide, a significant amount of methane was also being consistently released from the pipeline as a result of leaks and breakdowns. In 2001, natural gas emissions from the pipeline and natural gas transportation system accounted for 1% of the natural gas produced.[60] Between 2001 and 2005, this was reduced to 0.7%, the 2001 value was significantly less than that of 1996.[60]

A 2012 Climatic Change article and 2014 publication by a team of scientists led by Robert W. Howarth said that there was strong evidence that "shale gas has a larger GHG footprint than conventional gas, considered over any time scale. The GHG footprint of shale gas also exceeds that of oil or coal when considered at decadal time scales."[61][62] Howarth called for policy changes to regulate methane emissions resulting from hydraulic fracturing and shale gas development.[63]

A 2013 study by a team of researchers led by Scot M. Miller, said that U.S. greenhouse gas reduction policies in 2013 were based on what appeared to be significant underestimates of anthropogenic methane emissions.[64] The article said, that "greenhouse gas emissions from agriculture and fossil fuel extraction and processing"—oil and/or natural gas—were "likely a factor of two or greater than cited in existing studies."[64] By 2001, following a detailed study anthropogenic sources on climate change, IPCC researchers found that there was "stronger evidence that most of the observed warming observed over the last 50 years [was] attributable to human activities."[65][66] Since the Industrial Revolution humans have had a major impact on concentrations of atmospheric methane, increasing atmospheric concentrations roughly 250%.[67] According to the 2021 IPCC report, 30 - 50% of the current rise in temperatures is caused by emissions of methane,[68] and reducing methane is a fast way of climate change mitigation.[69] An alliance of 107 countries, including Brazil, the EU and the US, have joined the pact known as the Global Methane Pledge, committing to a collective goal of reducing global methane emissions by at least 30% from 2020 levels by 2030.[70][71]

Animals and livestock

Ruminant animals, particularly cows and sheep, contain bacteria in their gastrointestinal systems that help to break down plant material. Some of these microorganisms use the acetate from the plant material to produce methane, and because these bacteria live in the stomachs and intestines of ruminants, whenever the animal "burps" or defecates, it emits methane as well. Based upon a 2012 study in the Snowy Mountains region, the amount of methane emitted by one cow is equivalent to the amount of methane that around 3.4 hectares of methanotrophic bacteria can consume.[72]: 103  research in the Snowy Mountains region of Australia showed 8 tonnes of methane oxidized by methanotrophic bacteria per year on a 1,000 hectare farm. 200 cows on the same farm emitted 5.4 tonnes of methane per year. Hence, one cow emitted 27 kg of methane per year, while the bacteria oxidized 8 kg per hectare. The emissions of one cow were oxidized by 27/8 ≈ 3.4 hectare.

Termites also contain methanogenic microorganisms in their gut. However, some of these microorganisms are so unique that they live nowhere else in the world except in the third gut of termites. These microorganisms also break down biotic components to produce ethanol, as well as methane byproduct. However, unlike ruminants who lose 20% of the energy from the plants they eat, termites only lose 2% of their energy in the process.[73] Thus comparatively, termites do not have to eat as much food as ruminants to obtain the same amount of energy, and give off proportionally less methane.

In 2001, NASA researchers confirmed the vital role of enteric fermentation in livestock on global warming.[74] A 2006 UN FAO report reported that livestock generate more greenhouse gases as measured in CO2 equivalents than the entire transportation sector. Livestock accounts for 9% of anthropogenic CO2, 65%t of anthropogenic nitrous oxide and 37% of anthropogenic methane.[75] Since then, animal science and biotechnology researchers have focused research on methanogens in the rumen of livestock and mitigation of methane emissions.[76]

Nicholas Stern, the author of the 2006 Stern Review on climate change has stated "people will need to turn vegetarian if the world is to conquer climate change".[77] In 2003, the National Academy of Sciences's president, Ralph Cicerone—an atmospheric scientist—raised concerns about the increase in the number of methane-producing dairy and beef cattle was a "serious topic" as methane was the "second-most-important greenhouse gas in the atmosphere".[78]

Approximately 5% of the methane is released via the flatus, whereas the other 95% is released via eructation. Vaccines are under development to reduce the amount introduced through eructation.[79] Asparagopsis seaweed as a livestock feed additive has reduced methane emissions by more than 80%.[80]

Waste

Landfills

Due to the large collections of organic matter and availability of anaerobic conditions, landfills are the third largest source of atmospheric methane in the United States, accounting for roughly 18.2% of methane emissions globally in 2014.[81] When waste is first added to a landfill, oxygen is abundant and thus undergoes aerobic decomposition; during which time very little methane is produced. However, generally within a year oxygen levels are depleted and anaerobic conditions dominate the landfill allowing methanogens to takeover the decomposition process. These methanogens emit methane into the atmosphere and even after the landfill is closed, the mass amount of decaying matter allows the methanogens to continue producing methane for years.[82]

Waste water treatment

Waste water treatment facilities act to remove organic matter, solids, pathogens, and chemical hazards as a result of human contamination. Methane emission in waste treatment facilities occurs as a result of anaerobic treatments of organic compounds and anaerobic biodegradation of sludge.[83]

Others

Aquatic ecosystems

Natural and anthropogenic methane emissions from aquatic ecosystems are estimated to contribute about half of total global emissions.[84] Urbanization and eutrophication are expected to lead to increased methane emissions from aquatic ecosystems.[84]

Ecological conversion

Conversion of forests and natural environments into agricultural plots increases the amount of nitrogen in the soil, which inhibits methane oxidation, weakening the ability of the methanotrophic bacteria in the soil to act as sinks.[85] Additionally, by changing the level of the water table, humans can directly affect the soil's ability to act as a source or sink. The relationship between water table levels and methane emission is explained in the wetlands section of natural sources.

Rice agriculture

Rice agriculture is a significant source of methane. With warm weather and water-logged soil, rice paddies act like wetlands, but are generated by humans for the purpose of food production. Due to the swamp-like environment of rice fields, these paddies emitted about 30 of the 400 million metric tons of anthropogenic methane in 2022.[86]

Biomass burning

Incomplete burning of both living and dead organic matter results in the emission of methane. While natural wildfires can contribute to methane emissions, the bulk majority of biomass burning occurs as a result of humans – including everything from accidental burnings by civilians to deliberate burnings used to clear out land to biomass burnings occurring as a result of destroying waste.[87]

Oil and natural gas supply chain

Methane is a primary component of natural gas, and thus during the production, processing, storage, transmission, and distribution of natural gas, a significant amount of methane is lost into the atmosphere.[83]

According to the EPA Inventory of U.S Greenhouse Gas Emissions and Sinks: 1990–2015 report, 2015 methane emissions from natural gas and petroleum systems totaled 8.1 Tg per year in the United States. Individually, the EPA estimates that the natural gas system emitted 6.5 Tg per year of methane while petroleum systems emitted 1.6 Tg per year of methane.[88] Methane emissions occur in all sectors of the natural gas industry, from drilling and production, through gathering and processing and transmission, to distribution. These emissions occur through normal operation, routine maintenance, fugitive leaks, system upsets, and venting of equipment. In the oil industry, some underground crude contains natural gas that is entrained in the oil at high reservoir pressures. When oil is removed from the reservoir, associated gas is produced.

However, a review of methane emissions studies reveals that the EPA Inventory of Greenhouse Gas Emissions and Sinks: 1990–2015 report likely significantly underestimated 2015 methane emissions from the oil and natural gas supply chain. The review concluded that in 2015 the oil and natural gas supply chain emitted 13 Tg per year of methane, which is about 60% more than the EPA report for the same time period. The authors write that the most likely cause for the discrepancy is an under sampling by the EPA of so-called "abnormal operating conditions", during which large quantities of methane can be emitted.[89]

2015 methane emissions from oil and natural gas supply chain in the United States (Tg per year)
Supply chain segment EPA Inventory of US Greenhouse Gas

Emissions and Sinks: 1990–2015 report[88]

Alvarez et al. 2018[89]
Oil and natural gas production 3.5 7.6
Natural gas gathering 2.3 2.6
Natural gas transmission and storage 1.4 1.8
Natural gas processing 0.44 0.72
Natural gas local distribution 0.44 0.44
Oil refining and transportation 0.034 0.034
Total (95% confidence interval) 8.1 (6.7–10.2) 13 (11.3–15.1)

Coal mining

In 2014 NASA researchers reported the discovery of a 2,500 square miles (6,500 km2) methane cloud floating over the Four Corners region of the south-west United States. The discovery was based on data from the European Space Agency's Scanning Imaging Absorption Spectrometer for Atmospheric Chartography instrument from 2002 to 2012.[90]

The report concluded that "the source is likely from established gas, coal, and coalbed methane mining and processing." The region emitted 590,000 metric tons of methane every year between 2002 and 2012—almost 3.5 times the widely used estimates in the European Union's Emissions Database for Global Atmospheric Research.[90] In 2019, the International Energy Agency (IEA) estimated that the methane emissions leaking from the world's coalmines are warming the global climate at the same rate as the shipping and aviation industries combined.[91]

Permafrost thawing

Image showing melted permafrost resulting in thermokarst, a source of methane released from permafrost.

Permafrost contains almost twice as much carbon as the atmosphere,[92] with ~20 Gt of permafrost-associated methane trapped in methane clathrates.[93] Permafrost thaw results in the formation of thermokarst lakes in ice-rich yedoma deposits.[94] Methane frozen in permafrost is slowly released as permafrost melts.[95] Radiocarbon dating of trace methane in lake bubbles and soil organic carbon concluded that 0.2 to 2.5 Pg of permafrost carbon has been released as methane and carbon dioxide over the last 60 years.[96] The 2020 heat wave may have released significant methane from carbonate deposits in Siberian permafrost.[97]

Methane emissions by the 'permafrost carbon feedback' -- amplification of surface warming due to enhanced radiative forcing by carbon release from permafrost—could contribute an estimated 205 Gt of carbon emissions, leading up to 0.5 °C (0.9 °F) of additional warming by the end of the 21st century.[98] However, recent research based on the carbon isotopic composition of atmospheric methane trapped in bubbles in Antarctic ice suggests that methane emissions from permafrost and methane hydrates were minor during the last deglaciation, suggesting that future permafrost methane emissions may be lower than previously estimated.[99]

Methane gas from methane clathrates

Arctic methane concentrations up to September 2020.

At high pressures, such as are found on the bottom of the ocean, methane forms a solid clathrate with water, known as methane hydrate. An unknown, but possibly very large quantity of methane is trapped in this form in ocean sediments.

Theories suggest that should global warming cause them to heat up sufficiently, all of this methane gas could again be released into the atmosphere. Since methane gas is twenty-five times stronger (for a given weight, averaged over 100 years) than CO
2
as a greenhouse gas; this would immensely magnify the greenhouse effect.

The 2021 IPCC Sixth Assessment Report (AR6) Working Group 1 report said that it was "very unlikely that gas clathrates (mostly methane) in deeper terrestrial permafrost and subsea clathrates will lead to a detectable departure from the emissions trajectory during this century".[51]: 5 

Methane slip from gas engines

The use of natural gas and biogas in internal combustion engines for such applications as electricity production, cogeneration and heavy vehicles or marine vessels such as LNG carriers using the boil off gas for propulsion, emits a certain percentage of unburned hydrocarbons of which 85% is methane. The climate issues of using gas to fuel internal combustion engines may offset or even cancel out the advantages of less CO2 and particle emissions is described in this 2016 EU Issue Paper on methane slip from marine engines: "Emissions of unburnt methane (known as the 'methane slip') were around 7 g per kg LNG at higher engine loads, rising to 23–36 g at lower loads. This increase could be due to slow combustion at lower temperatures, which allows small quantities of gas to avoid the combustion process". Road vehicles run more on low load than marine engines causing relatively higher methane slip.

Release of stored arctic methane due to global warming

Global warming due to fossil fuel emissions has caused Arctic methane release, i.e. the release of methane from seas and soils in permafrost regions of the Arctic. Although in the long term, this is a natural process, methane release is being exacerbated and accelerated by global warming. This results in negative effects, as methane is itself a powerful greenhouse gas.

The Arctic region is one of the many natural sources of the greenhouse gas methane.[100] Global warming accelerates its release, due to both release of methane from existing stores, and from methanogenesis in rotting biomass.[101] Large quantities of methane are stored in the Arctic in natural gas deposits, permafrost, and as undersea clathrates. Permafrost and clathrates degrade on warming,[102] thus large releases of methane from these sources may arise as a result of global warming.[103][104][105] Other sources of methane include submarine taliks, river transport, ice complex retreat, submarine permafrost and decaying gas hydrate deposits.[106]

Global methane emissions monitoring

Methane (CH4) measured by the Advanced Global Atmospheric Gases Experiment (AGAGE) in the lower atmosphere (troposphere) at stations around the world. Abundances are given as pollution free monthly mean mole fractions in parts-per-billion

The Tropospheric Monitoring Instrument aboard the European Space Agency's Sentinel-5P spacecraft launched in October 2017 provides the most detailed methane emissions monitoring which is publicly available. It has a resolution of about 50 square kilometres.[107]

MethaneSAT is under development by the Environmental Defense Fund in partnership with researchers at Harvard University, to monitor methane emissions with an improved resolution of 1 kilometer. MethaneSAT is designed to monitor 50 major oil and gas facilities, and could also be used for monitoring of landfills and agriculture. It receives funding from Audacious Project (a collaboration of TED and the Gates Foundation), and is projected to launch as soon as 2024.[108]

Uncertainties in methane emissions, including so-called "super-emitter" fossil extractions[109] and unexplained atmospheric fluctuations,[110] highlight the need for improved monitoring at both regional and global scale. Satellites have recently begun to come online with capability to measure methane and other more powerful greenhouse gases with improving resolution.[111][112][113]

The Tropomi[114] instrument on Sentinel-5 launched in 2017 by the European Space Agency can measure methane, sulphur dioxide, nitrogen dioxide, carbon monoxide, aerosol, and ozone concentrations in earth's troposphere at resolutions of several kilometers.[109][115][116] In 2022, a study using data from the instrument monitoring large methane emissions worldwide was published; 1,200 large methane plumes were detected over oil and gas extraction sites.[117] NASA's EMIT instrument also identified super-emitters.[118]

Japan's GOSAT-2 platform launched in 2018 provides similar capability.[119]

The Claire satellite launched in 2016 by the Canadian firm GHGSat uses data from Tropomi to home in on sources of methane emissions as small as 15 m2.[111]

Other satellites are planned that will increase the precision and frequency of methane measurements, as well as provide a greater ability to attribute emissions to terrestrial sources. These include MethaneSAT, expected to be launched in 2022, and CarbonMapper.

Global maps combining satellite data to help identify and monitor major methane emission sources are being built.[120][121][122]

The International Methane Emissions Observatory was created by the UN.

Quantifying the global methane budget

In order to mitigate climate change, scientists have been focusing on quantifying the global methane CH4 budget as the concentration of methane continues to increase—it is now second after carbon dioxide in terms of climate forcing.[123] Further understanding of atmospheric methane is necessary in "assessing realistic pathways" towards climate change mitigation.[123] Various research groups give the following values for methane emissions:

Estimates of the global methane budget (in Tg(CH
4
)/yr)
Reference: Fung et al. (1991)[124] Hein et al. (1997)[124] Lelieveld et al. (1998)[124] Houweling et al. (1999)[124] Bousquet et al. (2006)[125] Saunois et al. (2016)[126] Saunois et al. (2020)[127]
Base year: 1980s 1992 2003–2012 2008-2017
Natural sources
Wetlands 115 237 225[nb 1] 145 147±15 167 (127–202) 181 (159-200)
Termites 20 20 20 23±4 64 (21–132) 37 (21–50)
Ocean 10 15 15 19±6
Hydrates 5 10
Anthropogenic sources
Energy 75 97 110 89 110±13 105 (77–133) 111 (81-131)
Landfills 40 35 40 73 55±11[nb 2] 188 (115-243) 217 (207-240)
Ruminants (livestock) 80 90[nb 3] 115 93
Waste treatment [nb 3] 25 [nb 2]
Rice agriculture 100 88 [nb 1] 31±5
Biomass burning 55 40 40 50±8 34 (15–53) 30 (22-36)
Other 20 90±14[nb 4]
Sinks
Soils 10 30 40 21±3 33 (28–38) 38 (27-45)
Tropospheric OH 450 489 510 448±1 515 518 (474–532)
Stratospheric loss 46 40 37±1
Source versus sink imbalance
Total source 500 587 600 525±8 558 (540–568) 576 (550-594)
Total sink 460 535 580 506 548 556 (501–574)

National reduction policies

An International Energy Agency graphic showing the potential of various emission reduction policies for addressing global methane emissions.
Global anthropogenic methane emissions from historical inventories and future Shared Socioeconomic Pathways (SSP) projections.[17]

China implemented regulations requiring coal plants to either capture methane emissions or convert methane into CO2 in 2010. According to a Nature Communications paper published in January 2019, methane emissions instead increased 50 percent between 2000 and 2015.[128][129]

In March 2020, Exxon called for stricter methane regulations, which would include detection and repair of methane leaks, minimization of venting and releases of unburned methane, and reporting requirements for companies.[130] However, in August 2020, the U.S. Environmental Protection Agency rescinded a prior tightening of methane emission rules for the U.S. oil and gas industry.[131][132]

Methane emissions for 2017 by region, source category, and latitude.[133]

Approaches to reduce emissions

Natural gas industries

About 40% of methane emissions from the fossil fuel industry could be "eliminated at no net cost for firms", according to the International Energy Agency (IEA) by using existing technologies.[15] Forty percent represents 9% of all human methane emissions.[15]

To reduce emissions from the natural gas industries, the EPA developed the Natural Gas STAR Program, also known as Gas STAR.[83]

The Coalbed Methane Outreach Program (CMOP) helps and encourages the mining industry to find ways to use or sell methane that would otherwise be released from the coal mine into the atmosphere.[83]

In 2023, the European Union agreed to legislation that will require fossil fuel companies to monitor and report methane leaks and to repair them within a short time period. The law also compels remediation of methane venting and  methane flaring. The United States and China stated that they will include methane reduction targets in their next climate plans but have not enacted rules that would compel monitoring, reporting or repair of methane leaks.[134]

Livestock

In order to counteract the amount of methane that ruminants give off, a type of drug called monensin (marketed as rumensin) has been developed. This drug is classified as an ionophore, which is an antibiotic that is naturally produced by a harmless bacteria strain. This drug not only improves feed efficiency but also reduces the amount of methane gas emitted from the animal and its manure.[135]

In addition to medicine, specific manure management techniques have been developed to counteract emissions from livestock manure. Educational resources have begun to be provided for small farms. Management techniques include daily pickup and storage of manure in a completely closed off storage facility that will prevent runoff from making it into bodies of water. The manure can then be kept in storage until it is either reused for fertilizer or taken away and stored in an offsite compost. Nutrient levels of various animal manures are provided for optimal use as compost for gardens and agriculture.[136]

Crops and soils

In order to reduce effects on methane oxidation in soil, several steps can be taken. Controlling the usage of nitrogen enhancing fertilizer and reducing the amount of nitrogen pollution into the air can both lower inhibition of methane oxidation. Additionally, using drier growing conditions for crops such as rice and selecting strains of crops that produce more food per unit area can reduce the amount of land with ideal conditions for methanogenesis. Careful selection of areas of land conversion (for example, plowing down forests to create agricultural fields) can also reduce the destruction of major areas of methane oxidation.[citation needed]

Landfills

To counteract methane emissions from landfills, on March 12, 1996, the EPA (Environmental Protection Agency) added the "Landfill Rule" to the Clean Air Act. This rule requires large landfills that have ever accepted municipal solid waste, have been used as of November 8, 1987, can hold at least 2.5 million metric tons of waste with a volume greater than 2.5 million cubic meters, and/or have nonmethane organic compound (NMOC) emissions of at least 50 metric tons per year to collect and combust emitted landfill gas.[137] This set of requirements excludes 96% of the landfills in the USA. While the direct result of this is landfills reducing emission of non-methane compounds that form smog, the indirect result is reduction of methane emissions as well.

In an attempt to absorb the methane that is already being produced from landfills, experiments in which nutrients were added to the soil to allow methanotrophs to thrive have been conducted. These nutrient supplemented landfills have been shown to act as a small scale methane sink, allowing the abundance of methanotrophs to sponge the methane from the air to use as energy, effectively reducing the landfill's emissions.[138]

See also

Notes

  1. ^ a b Rice included under wetlands.
  2. ^ a b Landfills total includes domestic sewage and animal waste.
  3. ^ a b Waste treatment included under ruminants.
  4. ^ Contains a small amount of natural emissions from wild ruminants

References

  1. ^ a b c d e f g "Global Methane Emissions and Mitigation Opportunities" (PDF). Global Methane Initiative. 2020.
  2. ^ a b IPCC Fifth Assessment Report - Radiative Forcings (AR5 Figure SPM.5) (Report). Intergovernmental Panel on Climate Change. 2013.
  3. ^ a b c d "Sources of methane emissions". International Energy Agency. Retrieved 2020-08-20.
  4. ^ a b "Global Carbon Project (GCP)". www.globalcarbonproject.org. Retrieved 2019-07-25.
  5. ^ Methane - A compelling case for action (Report). International Energy Agency. 2020-08-20.
  6. ^ Borunda, A. (2021, May 03). Methane facts and information. Retrieved April 6, 2022, from [1]
  7. ^ Leber, Rebecca (2021-08-12). "It's time to freak out about methane emissions". Vox. Retrieved 2022-01-05.
  8. ^ Trakimavicius, Lukas. "Putting a lid on Russia's planet-heating methane emissions". EurActiv. Retrieved 2023-07-26.
  9. ^ Timothy Puko (19 October 2021). "Who Are the World's Biggest Climate Polluters? Satellites Sweep for Culprits". The Wall Street Journal. Retrieved 19 October 2021. Russia is the world's top source of methane emissions from the oil-and-gas industry
  10. ^ "Yes, cattle are the top source of methane emissions in the US". verifythis.com. 2021-11-12. Retrieved 2024-02-26.
  11. ^ "Cows and Climate Change". UC Davis. 2019-06-27. Retrieved 2024-02-26.
  12. ^ Johnson, K A (1995-08-01). "Methane emissions from cattle". academic.oup.com. Retrieved 2023-04-27.
  13. ^ a b "Methane, explained". National Geographic. nationalgeographic.com. 2019-01-23. Archived from the original on April 17, 2019. Retrieved 2019-07-25.
  14. ^ Vaclav Smil (2017-03-29). "Planet of the Cows". IEEE Spectrum. Retrieved 2020-09-08.
  15. ^ a b c "Governments should set targets to reduce methane emissions". The Economist. 2021-03-31. ISSN 0013-0613. Retrieved October 10, 2021.
  16. ^ Terazono, Emiko; Hodgson, Camilla (2021-10-10). "How methane-producing cows leapt to the frontline of climate change". Financial Times. Retrieved October 10, 2021.
  17. ^ a b c d e f Saunois, Marielle; Stavert, Ann R.; Poulter, Ben; Bousquet, Philippe; Canadell, Josep G.; Jackson, Robert B.; Raymond, Peter A.; Dlugokencky, Edward J.; Houweling, Sander; Patra, Prabir K.; Ciais, Philippe; Arora, Vivek K.; Bastviken, David; Bergamaschi, Peter; Blake, Donald R.; Brailsford, Gordon; Bruhwiler, Lori; Carlson, Kimberly M.; Carrol, Mark; Castaldi, Simona; Chandra, Naveen; Crevoisier, Cyril; Crill, Patrick M.; Covey, Kristofer; Curry, Charles L.; Etiope, Giuseppe; Frankenberg, Christian; Gedney, Nicola; Hegglin, Michaela I.; et al. (15 July 2020). "The Global Methane Budget 2000–2017". Earth System Science Data. 12 (3): 1561–1623. Bibcode:2020ESSD...12.1561S. doi:10.5194/essd-12-1561-2020. ISSN 1866-3508. Retrieved 28 August 2020.
  18. ^ "Increase in atmospheric methane set another record during 2021 / Carbon dioxide levels also record a big jump". NOAA.gov. 7 April 2022. Archived from the original on 7 April 2022.
  19. ^ Earth System Research Laboratory Global Monitoring Division, NOAA, May 5, 2019
  20. ^ Butler J. and Montzka S. (2020). "The NOAA Annual Greenhouse Gas Index (AGGI)". NOAA Global Monitoring Laboratory/Earth System Research Laboratories.
  21. ^ Boucher O, Friedlingstein P, Collins B, Shine KP (2009). "The indirect global warming potential and global temperature change potential due to methane oxidation". Environ. Res. Lett. 4 (4): 044007. Bibcode:2009ERL.....4d4007B. doi:10.1088/1748-9326/4/4/044007.
  22. ^ "Understanding Global Warming Potentials". 12 January 2016. Retrieved 2019-09-09.
  23. ^ a b Myhre, G., D. Shindell, F.-M. Bréon, W. Collins, J. Fuglestvedt, J. Huang, D. Koch, J.-F. Lamarque, D. Lee, B. Mendoza, T. Nakajima, A. Robock, G. Stephens, T. Takemura and H. Zhang (2013) "Anthropogenic and Natural Radiative Forcing". Table 8.7 on page 714. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.). Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Anthropogenic and Natural Radiative Forcing
  24. ^ Terazono, Emiko; Hodgson, Camilla (2021-10-10). "How methane-producing cows leapt to the frontline of climate change". Financial Times. Retrieved 2021-10-10.
  25. ^ "Governments should set targets to reduce methane emissions". The Economist. 2021-03-31. ISSN 0013-0613. Retrieved 2021-10-10.
  26. ^ Schwietzke, S., Sherwood, O., Bruhwiler, L.; et al. (2016). "Upward revision of global fossil fuel methane emissions based on isotope database". Nature. Springer Nature. 538 (7623): 88–91. Bibcode:2016Natur.538...88S. doi:10.1038/nature19797. PMID 27708291. S2CID 4451521.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  27. ^ Hmiel, B., Petrenko, V.V., Dyonisius, M.N.; et al. (2020). "Preindustrial 14CH4 indicates greater anthropogenic fossil CH4 emissions". Nature. Springer Nature. 578 (7795): 409–412. Bibcode:2020Natur.578..409H. doi:10.1038/s41586-020-1991-8. PMID 32076219. S2CID 211194542.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  28. ^ Adam Voiland and Joshua Stevens (8 March 2016). "Methane Matters". NASA Earth Observatory. Retrieved 2020-09-15.
  29. ^ "CH
    4
     responsible for more than 80% of recent atmospheric methane growth"
    . UPI. Retrieved 27 April 2022.
  30. ^ Feng, Liang; Palmer, Paul I.; Zhu, Sihong; Parker, Robert J.; Liu, Yi (16 March 2022). "Tropical methane emissions explain large fraction of recent changes in global atmospheric methane growth rate". Nature Communications. 13 (1): 1378. Bibcode:2022NatCo..13.1378F. doi:10.1038/s41467-022-28989-z. ISSN 2041-1723. PMC 8927109. PMID 35297408.
  31. ^ "Gas flares aren't as efficient at burning off methane as assumed". Science News. 29 September 2022. Retrieved 21 October 2022.
  32. ^ Plant, Genevieve; Kort, Eric A.; Brandt, Adam R.; Chen, Yuanlei; Fordice, Graham; Gorchov Negron, Alan M.; Schwietzke, Stefan; Smith, Mackenzie; Zavala-Araiza, Daniel (30 September 2022). "Inefficient and unlit natural gas flares both emit large quantities of methane". Science. 377 (6614): 1566–1571. Bibcode:2022Sci...377.1566P. doi:10.1126/science.abq0385. ISSN 0036-8075. PMID 36173866. S2CID 252621958.
  33. ^ Hmiel, Benjamin; Petrenko, V. V.; Dyonisius, M. N.; Buizert, C.; Smith, A. M.; Place, P. F.; Harth, C.; Beaudette, R.; Hua, Q.; Yang, B.; Vimont, I.; Michel, S. E.; Severinghaus, J. P.; Etheridge, D.; Bromley, T.; Schmitt, J.; Faïn, X.; Weiss, R. F.; Dlugokencky, E. (20 February 2020). "Preindustrial 14CH4 indicates greater anthropogenic fossil CH4 emissions". Nature. 578 (7795): 409–412. Bibcode:2020Natur.578..409H. doi:10.1038/s41586-020-1991-8. PMID 32076219. S2CID 211194542.
  34. ^ Gorchov Negron, Alan M.; Kort, Eric A.; Conley, Stephen A.; Smith, Mackenzie L. (21 April 2020). "Airborne Assessment of Methane Emissions from Offshore Platforms in the U.S. Gulf of Mexico". Environmental Science & Technology. 54 (8): 5112–5120. Bibcode:2020EnST...54.5112G. doi:10.1021/acs.est.0c00179. ISSN 0013-936X. PMID 32281379.
  35. ^ Zhang, Yuzhong; Gautam, Ritesh; Pandey, Sudhanshu; Omara, Mark; Maasakkers, Joannes D.; Sadavarte, Pankaj; Lyon, David; Nesser, Hannah; Sulprizio, Melissa P.; Varon, Daniel J.; Zhang, Ruixiong; Houweling, Sander; Zavala-Araiza, Daniel; Alvarez, Ramon A.; Lorente, Alba; Hamburg, Steven P.; Aben, Ilse; Jacob, Daniel J. (1 April 2020). "Quantifying methane emissions from the largest oil-producing basin in the United States from space". Science Advances. 6 (17): eaaz5120. Bibcode:2020SciA....6.5120Z. doi:10.1126/sciadv.aaz5120. PMC 7176423. PMID 32494644.
  36. ^ "Zero Routine Flaring by 2030". World Bank. Retrieved 2020-09-18.
  37. ^ a b "About Methane". Global Methane Initiative. Retrieved 2020-09-15.
  38. ^ a b US EPA, OA (23 December 2015). "Overview of Greenhouse Gases". US EPA.
  39. ^ "Agriculture's greenhouse gas emissions on the rise". FAO. Retrieved 2017-04-19.
  40. ^ "Fossil fuel industry's methane emissions far higher than thought". The Guardian. 2016. Emissions of the powerful greenhouse gas from coal, oil and gas are up to 60% greater balls than previously estimated, meaning current climate prediction models should be revised, research shows
  41. ^ Bruhn, D.; et al. (March 2014). "Leaf surface wax is a source of plant methane formation under UV radiation and in the presence of oxygen". Plant Biology. 16 (2): 512–516. doi:10.1111/plb.12137. PMID 24400835.
  42. ^ Volodin, E. M. (May 2015). "Influence of methane sources in Northern Hemisphere high latitudes on the interhemispheric asymmetry of its atmospheric concentration and climate". Izvestiya, Atmospheric and Oceanic Physics. 51 (3): 251–258. Bibcode:2015IzAOP..51..251V. doi:10.1134/S0001433815030123. S2CID 118933772.
  43. ^ Crevoisier, C.; et al. (September 2012). "The 2007–2011 evolution of tropical methane in the mid-troposphere as seen from space by MetOp-A/IASI" (PDF). Atmospheric Chemistry and Physics Discussions. 12 (9): 23731–23757. Bibcode:2013ACP....13.4279C. doi:10.5194/acpd-12-23731-2012.
  44. ^ a b Bubier, Jill L.; Moore, Tim R. (December 1994). "An ecological perspective on methane emissions from northern wetlands". Trends in Ecology and Evolution. 9 (12): 460–464. doi:10.1016/0169-5347(94)90309-3. PMID 21236923.
  45. ^ Houghton, J. T., et al. (Eds.) (2001) Projections of future climate change, Climate Change 2001: The Scientific Basis, Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change, 881 pp.
  46. ^ Comyn-Platt, Edward (2018). "Carbon budgets for 1.5 and 2 °C targets lowered by natural wetland and permafrost feedbacks" (PDF). Nature. 11 (8): 568–573. Bibcode:2018NatGe..11..568C. doi:10.1038/s41561-018-0174-9. S2CID 134078252.
  47. ^ Bridgham, Scott D.; Cadillo-Quiroz, Hinsby; Keller, Jason K.; Zhuang, Qianlai (May 2013). "Methane emissions from wetlands: biogeochemical, microbial, and modeling perspectives from local to global scales". Global Change Biology. 19 (5): 1325–1346. Bibcode:2013GCBio..19.1325B. doi:10.1111/gcb.12131. PMID 23505021. S2CID 14228726.
  48. ^ Saunois, Marielle; Stavert, Ann R.; Poulter, Ben; Bousquet, Philippe; Canadell, Josep G.; Jackson, Robert B.; Raymond, Peter A.; Dlugokencky, Edward J.; Houweling, Sander; Patra, Prabir K.; Ciais, Philippe; Arora, Vivek K.; Bastviken, David; Bergamaschi, Peter; Blake, Donald R. (2020-07-15). "The Global Methane Budget 2000–2017". Earth System Science Data. 12 (3): 1561–1623. doi:10.5194/essd-12-1561-2020. ISSN 1866-3508.
  49. ^ Christensen, T. R., A. Ekberg, L. Strom, M. Mastepanov, N. Panikov, M. Oquist, B. H. Svenson, H. Nykanen, P. J. Martikainen, and H. Oskarsson (2003), Factors controlling large scale variations in methane emissions from wetlands, Geophys. Res. Lett., 30, 1414, doi:10.1029/2002GL016848.
  50. ^ Carmichael, J.; et al. (June 2014). "The role of vegetation in methane flux to the atmosphere: should vegetation be included as a distinct category in the global methane budget?". Biogeochemistry. 119 (1): 1–24. doi:10.1007/s10533-014-9974-1. S2CID 13533695.
  51. ^ a b Fox-Kemper, B.; Hewitt, H.T.; Xiao, C.; Aðalgeirsdóttir, G.; Drijfhout, S.S.; Edwards, T.L.; Golledge, N.R.; Hemer, M.; Kopp, R.E.; Krinner, G.; Mix, A. (2021). Masson-Delmotte, V.; Zhai, P.; Pirani, A.; Connors, S.L.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L. (eds.). "Chapter 5: Global Carbon and other Biogeochemical Cycles and Feedbacks". Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York, NY, USA. doi:10.1017/9781009157896.011. Archived from the original (PDF) on January 20, 2023.
  52. ^ Global Methane Assessment (PDF). United Nations Environment Programme and Climate and Clean Air Coalition (Report). Nairobi. 2022. p. 12. Retrieved March 15, 2023.
  53. ^ "Climate Change 2021. The Physical Science Basis. Summary for Policymakers. Working Group I contribution to the WGI Sixth Assessment Report of the Intergovernmental Panel on Climate Change". IPCC. The Intergovernmental Panel on Climate Change. Archived from the original on August 22, 2021. Retrieved August 22, 2021.
  54. ^ Shindell, Drew, ed. (May 6, 2021). Global Methane Assessment: Benefits and Costs of Mitigating Methane Emissions. United Nations Environment Programme (Report). p. 173. ISBN 978-92-807-3854-4.
  55. ^ "Ch.2 Changes in Atmospheric Constituents and in Radiative Forcing". Climate Change 2007 IPCC Fourth Assessment Report. IPPC. Retrieved 2017-01-20.
  56. ^ Nisbet, E. G.; Manning, M. R.; Dlugokencky, E. J.; Fisher, R. E.; Lowry, D.; Michel, S. E.; Myhre, C. Lund; Platt, S. M.; Allen, G.; Bousquet, P.; Brownlow, R.; Cain, M.; France, J. L.; Hermansen, O.; Hossaini, R.; Jones, A. E.; Levin, I.; Manning, A. C.; Myhre, G.; Pyle, J. A.; Vaughn, B. H.; Warwick, N. J.; White, J. W. C. (2019). "Very Strong Atmospheric Methane Growth in the 4 Years 2014–2017: Implications for the Paris Agreement". Global Biogeochemical Cycles. 33 (3): 318–342. Bibcode:2019GBioC..33..318N. doi:10.1029/2018GB006009. ISSN 1944-9224. S2CID 133716021.
  57. ^ McKie, Robin (February 17, 2019). "Sharp rise in methane levels threatens world climate targets". The Observer. ISSN 0029-7712. Retrieved March 17, 2023.
  58. ^ "Fracking boom tied to methane spike in the atmosphere". National Geographic. 2019-08-15. Archived from the original on August 18, 2019. Retrieved 2019-08-20.
  59. ^ "Primer on Short-Lived Climate Pollutants". Climate & Clean Air Coalition. Retrieved March 19, 2023.
  60. ^ a b c Lechtenböhmer, Stephan; et al. (2005). "Greenhouse Gas Emissions from the Russian Natural Gas Export Pipeline System" (PDF). Wuppertal Institute for Climate, Environment and Energy. Archived from the original (PDF) on 2012-03-14. Retrieved 2016-12-31.
  61. ^ Howarth, Robert W.; Santoro, Renee; Ingraffea, Anthony (January 10, 2012). "Venting and leaking of methane from shale gas development: response to Cathles et al" (PDF). Climatic Change. 113 (2): 537–549. Bibcode:2012ClCh..113..537H. doi:10.1007/s10584-012-0401-0. S2CID 154324540. Retrieved 2016-12-22.
  62. ^ Howarth, Robert W. (June 1, 2014). "A bridge to nowhere: methane emissions and the greenhouse gas footprint of natural gas". Energy Sci Eng. 2 (2): 47–60. doi:10.1002/ese3.35. hdl:1813/60821.
  63. ^ Howarth, Robert (October 8, 2015). "Methane emissions and climatic warming risk from hydraulic fracturing and shale gas development: implications for policy". Energy and Emission Control Technologies. 3: 45. doi:10.2147/EECT.S61539.
  64. ^ a b Miller, Scot M.; Wofsy, Steven C.; Michalak, Anna M.; Kort, Eric A.; Andrews, Arlyn E.; Biraud, Sebastien C.; Dlugokencky, Edward J.; Eluszkiewicz, Janusz; Fischer, Marc L.; Janssens-Maenhout, Greet; Miller, Ben R.; Miller, John B.; Montzka, Stephen A.; Nehrkorn, Thomas; Sweeney, Colm (December 10, 2013). "Anthropogenic emissions of methane in the United States". PNAS. 110 (50): 20018–20022. Bibcode:2013PNAS..11020018M. doi:10.1073/pnas.1314392110. PMC 3864315. PMID 24277804.
  65. ^ Houghton, J.E.T.; Ding, Y.; Griggs, David; Noguer, Maria; van der Linden, Paul; Dai, X.; Maskell, M.; Johnson, C. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change (IPCC). Cambridge: Cambridge University Press. p. 881. ISBN 978-0521807678. OCLC 46634335.
  66. ^ "Technical summary". Climate Change 2001. United Nations Environment Programme. Archived from the original on June 4, 2011. Retrieved June 4, 2009.
  67. ^ Mitchell, Logan; et al. (November 2013). "Constraints on the Late Holocene Anthropogenic Contribution to the Atmospheric Methane Budget". Science. 342 (6161): 964–966. Bibcode:2013Sci...342..964M. doi:10.1126/science.1238920. PMID 24264988. S2CID 39963336.
  68. ^ McGrath, Matt (2021-08-11). "Climate change: Curbing methane emissions will 'buy us time'". BBC News. Retrieved 2021-08-11.
  69. ^ Ramirez, Rachel (August 11, 2021). "Scientists say this invisible gas could seal our fate on climate change". CNN. Archived from the original on August 11, 2021. Retrieved 2021-08-11.
  70. ^ "Joint EU-US Press Release on the Global Methane Pledge". European Commission. 2021-09-18. Archived from the original on June 21, 2019. Retrieved 2021-11-02.
  71. ^ Wintour, Patrick (2021-11-02). "Biden to unveil pledge to slash global methane emissions by 30%". The Guardian. Archived from the original on November 2, 2021. Retrieved 2021-11-02.
  72. ^ Mason-Jones, David (2012). Should Meat be on the Menu?. Momentum. ISBN 978-1743340608.
  73. ^ Margonelli, Lisa (September 2008). "Gut Reactions". The Atlantic. Retrieved January 16, 2012.
  74. ^ "Methane Explosion Warmed The Prehistoric Earth, Possible Again". NASA/Goddard Space Flight Center, EOS Project Science Office (Press release). December 12, 2001. Retrieved March 22, 2023 – via ScienceDaily.
  75. ^ "Livestock a major threat to environment". United Nations Food and Agriculture Organization. November 29, 2006. Archived from the original on March 28, 2008. Retrieved November 4, 2011.
  76. ^ Patra, Amlan; Park, Tansol; Kim, Minseok; Yu, Zhongtang (January 26, 2017). "Rumen methanogens and mitigation of methane emission by anti-methanogenic compounds and substances". Journal of Animal Science and Biotechnology. 8 (1): 13. doi:10.1186/s40104-017-0145-9. ISSN 2049-1891. PMC 5270371. PMID 28149512.
  77. ^ Pagnamenta, Robin (2009-10-27). "Climate chief Lord Stern give up meat to save the planet". The Times. London.
  78. ^ Gary Polakovic (June 7, 2003). "Getting the Cows to Cool It". The Los Angeles Times. Retrieved November 4, 2011.
  79. ^ Rachel Nowak (September 25, 2004). "Burp vaccine cuts greenhouse gas emissions". New Scientist. Retrieved November 4, 2011.
  80. ^ "New company to reduce cows' methane using feed additive made from the seaweed". The Cattle Site. 2020-09-22.
  81. ^ "Greenhouse Gas Emissions". United States Environmental Protection Agency. Retrieved March 21, 2013.
  82. ^ Themelis, Nickolas J.; Ulloa, Priscilla A. (June 2007). "Methane generation in landfills". Renewable Energy. 32 (7): 1243–1257. doi:10.1016/j.renene.2006.04.020. Retrieved 2016-12-31.
  83. ^ a b c d "Sources and Emissions". US Environmental Protection Agency. July 12, 2006. Archived from the original on July 12, 2006. Retrieved 2017-01-20.
  84. ^ a b Rosentreter, Judith A.; Borges, Alberto V.; Deemer, Bridget R.; Holgerson, Meredith A.; Liu, Shaoda; Song, Chunlin; Melack, John; Raymond, Peter A.; Duarte, Carlos M.; Allen, George H.; Olefeldt, David (2021). "Half of global methane emissions come from highly variable aquatic ecosystem sources". Nature Geoscience. 14 (4): 225–230. Bibcode:2021NatGe..14..225R. doi:10.1038/s41561-021-00715-2. hdl:10754/668712. ISSN 1752-0908. S2CID 233030781.
  85. ^ Nazaries, Loïc; et al. (September 2013). "Methane, microbes and models: fundamental understanding of the soil methane cycle for future predictions". Environmental Microbiology. 15 (9): 2395–2417. doi:10.1111/1462-2920.12149. PMID 23718889.
  86. ^ "Country Inventory - Climate TRACE". climatetrace.org. Retrieved 2023-12-22.
  87. ^ "Methane and Nitrous Oxide Emissions From Natural Sources" (PDF). USA Environmental Protection Agency Office of Atmospheric Programs. April 2010. Archived from the original (PDF) on 2012-12-02. Retrieved 2017-01-20.
  88. ^ a b "Inventory of U.S. Greenhouse Gas Emissions and Sinks: 1990–2015" (PDF).
  89. ^ a b Alvarez, Ramón A.; Zavala-Araiza, Daniel; Lyon, David R.; Allen, David T.; Barkley, Zachary R.; Brandt, Adam R.; Davis, Kenneth J.; Herndon, Scott C.; Jacob, Daniel J. (2018-07-13). "Assessment of methane emissions from the U.S. oil and gas supply chain". Science. 361 (6398): 186–188. Bibcode:2018Sci...361..186A. doi:10.1126/science.aar7204. ISSN 0036-8075. PMC 6223263. PMID 29930092.
  90. ^ a b Gass, Henry (October 10, 2014). "How scientists overlooked a 2,500-square-mile cloud of methane over the Southwest". Christian Science Monitor. Retrieved October 24, 2014.
  91. ^ Ambrose, Jillian (2019-11-15). "Methane emissions from coalmines could stoke climate crisis – study". The Guardian. ISSN 0261-3077. Retrieved 2019-11-15.
  92. ^ Brouillette, Monique (2021). "How microbes in permafrost could trigger a massive carbon bomb". Nature. 591 (7850): 360–362. Bibcode:2021Natur.591..360B. doi:10.1038/d41586-021-00659-y. PMID 33731951. S2CID 232297719.
  93. ^ Ruppel, C. (2014). "Permafrost-Associated Gas Hydrate: Is It Really Approximately 1 % of the Global System?". Journal of Chemical & Engineering Data. 60 (2): 429–436. doi:10.1021/je500770m. ISSN 0021-9568.
  94. ^ Zandt, Michiel H.; Liebner, Susanne; Welte, Cornelia U. (2020). "Roles of Thermokarst Lakes in a Warming World". Trends in Microbiology. 28 (9): 769–779. doi:10.1016/j.tim.2020.04.002. ISSN 0966-842X. PMID 32362540. S2CID 218492291.
  95. ^ Intergovernmental Panel on Climate Change, "IPCC, 2021: Summary for Policymakers", Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge: Cambridge University Press
  96. ^ Walter Anthony, Katey; Daanen, Ronald; Anthony, Peter; Schneider von Deimling, Thomas; Ping, Chien-Lu; Chanton, Jeffrey P.; Grosse, Guido (2016). "Methane emissions proportional to permafrost carbon thawed in Arctic lakes since the 1950s". Nature Geoscience. 9 (9): 679–682. Bibcode:2016NatGe...9..679W. doi:10.1038/ngeo2795. ISSN 1752-0908. OSTI 1776496.
  97. ^ Froitzheim, Nikolaus; Majka, Jaroslaw; Zastrozhnov, Dmitry (2021). "Methane release from carbonate rock formations in the Siberian permafrost area during and after the 2020 heat wave". Proceedings of the National Academy of Sciences. 118 (32). Bibcode:2021PNAS..11807632F. doi:10.1073/pnas.2107632118. ISSN 0027-8424. PMC 8364203. PMID 34341110.
  98. ^ Schuur, E. a. G.; McGuire, A. D.; Schädel, C.; Grosse, G.; Harden, J. W.; Hayes, D. J.; Hugelius, G.; Koven, C. D.; Kuhry, P.; Lawrence, D. M.; Natali, S. M. (2015). "Climate change and the permafrost carbon feedback". Nature. 520 (7546): 171–179. Bibcode:2015Natur.520..171S. doi:10.1038/nature14338. ISSN 1476-4687. PMID 25855454. S2CID 4460926.
  99. ^ Dyonisius, M. N.; Petrenko, V. V.; Smith, A. M.; Hua, Q.; Yang, B.; Schmitt, J.; Beck, J.; Seth, B.; Bock, M.; Hmiel, B.; Vimont, I. (2020-02-21). "Old carbon reservoirs were not important in the deglacial methane budget". Science. 367 (6480): 907–910. Bibcode:2020Sci...367..907D. doi:10.1126/science.aax0504. ISSN 0036-8075. PMID 32079770. S2CID 211230350.
  100. ^ Bloom, A. A.; Palmer, P. I.; Fraser, A.; Reay, D. S.; Frankenberg, C. (2010). "Large-Scale Controls of Methanogenesis Inferred from Methane and Gravity Spaceborne Data" (PDF). Science. 327 (5963): 322–325. Bibcode:2010Sci...327..322B. doi:10.1126/science.1175176. PMID 20075250. S2CID 28268515.
  101. ^ Walter, K. M.; Chanton, J. P.; Chapin, F. S.; Schuur, E. A. G.; Zimov, S. A. (2008). "Methane production and bubble emissions from arctic lakes: Isotopic implications for source pathways and ages". Journal of Geophysical Research. 113 (G3): G00A08. Bibcode:2008JGRG..113.0A08W. doi:10.1029/2007JG000569.
  102. ^ Carrington, Damian (July 21, 2020). "First active leak of sea-bed methane discovered in Antarctica". The Guardian.
  103. ^ Zimov, Sa; Schuur, Ea; Chapin, Fs 3Rd (June 2006). "Climate change. Permafrost and the global carbon budget". Science. 312 (5780): 1612–3. doi:10.1126/science.1128908. ISSN 0036-8075. PMID 16778046. S2CID 129667039.{{cite journal}}: CS1 maint: numeric names: authors list (link)
  104. ^ Shakhova, Natalia (2005). "The distribution of methane on the Siberian Arctic shelves: Implications for the marine methane cycle". Geophysical Research Letters. 32 (9): L09601. Bibcode:2005GeoRL..32.9601S. doi:10.1029/2005GL022751.
  105. ^ "Scientists shocked by Arctic permafrost thawing 70 years sooner than predicted". The Guardian. Reuters. June 18, 2019. ISSN 0261-3077. Retrieved 2019-07-14.
  106. ^ Shakhova, Natalia; Semiletov, Igor (2007). "Methane release and coastal environment in the East Siberian Arctic shelf". Journal of Marine Systems. 66 (1–4): 227–243. Bibcode:2007JMS....66..227S. CiteSeerX 10.1.1.371.4677. doi:10.1016/j.jmarsys.2006.06.006.
  107. ^ Tollefson, Jeff (2018-04-11). "US environmental group wins millions to develop methane-monitoring satellite". Nature. 556 (7701): 283. Bibcode:2018Natur.556..283T. doi:10.1038/d41586-018-04478-6. PMID 29666485.
  108. ^ Powell, Alvin (24 March 2023). "Buying crucial time in climate change fight". The Harvard Gazette. Retrieved 27 March 2023.
  109. ^ a b Hiroko Tabuchi (2019-12-16). "A Methane Leak, Seen From Space, Proves to Be Far Larger Than Thought". New York Times.
  110. ^ E Roston and NS Malik (2020-04-06). "Methane emissions hit a new record and scientists can't say why". Bloomberg.com. Bloomberg News.
  111. ^ a b John Fialka (2018-03-09). "Meet the satellite that can pinpoint methane and carbon dioxide leaks". Scientific American.
  112. ^ "MethaneSAT". methanesat.org. Retrieved 2020-09-10.
  113. ^ Katz, Cheryl (2021-06-15). "In Push to Find Methane Leaks, Satellites Gear Up for the Hunt". Yale E360. Retrieved 2022-01-02.
  114. ^ "Tropomi". European Space Agency. Retrieved 2020-09-10.
  115. ^ Michelle Lewis (2019-12-18). "New satellite technology reveals Ohio gas leak released 60K tons of methane". Electrek.
  116. ^ Joost A de Gouw; et al. (2020). "Daily Satellite Observations of Methane from Oil and Gas Production Regions in the United States". Scientific Reports. Springer Nature. 10 (10): 1379. Bibcode:2020NatSR..10.1379D. doi:10.1038/s41598-020-57678-4. PMC 6987228. PMID 31992727.
  117. ^ "Massive methane emissions by oil and gas industry detected from space | CNRS". www.cnrs.fr.
  118. ^ Wall, Mike (25 October 2022). "Methane 'super-emitters' on Earth spotted by space station experiment". Space.com. Retrieved 29 November 2022.
  119. ^ "Greenhouse gases Observing SATellite-2 "IBUKI-2" (GOSAT-2)". Japan Aerospace Exploration Agency. Retrieved 2020-10-21.
  120. ^ "Climate change: Satellites map huge methane plumes from oil and gas". BBC News. 4 February 2022. Retrieved 16 March 2022.
  121. ^ "Cracking down on methane 'ultra emitters' is a quick way to combat climate change, researchers find". Washington Post. Retrieved 16 March 2022.
  122. ^ Lauvaux, T.; Giron, C.; Mazzolini, M.; d’Aspremont, A.; Duren, R.; Cusworth, D.; Shindell, D.; Ciais, P. (4 February 2022). "Global assessment of oil and gas methane ultra-emitters". Science. 375 (6580): 557–561. arXiv:2105.06387. Bibcode:2022Sci...375..557L. doi:10.1126/science.abj4351. ISSN 0036-8075. PMID 35113691. S2CID 246530897.
  123. ^ a b Saunois, M; Jackson, B.; Bousquet, P.; Poulter, B.; Canadell, J G (2016). "The growing role of methane in anthropogenic climate change". Environmental Research Letters. Vol. 11, no. 120207. p. 120207. doi:10.1088/1748-9326/11/12/120207.
  124. ^ a b c d "Trace Gases: Current Observations, Trends, and Budgets". Climate Change 2001, IPCC Third Assessment Report. IPCC/United Nations Environment Programme. Archived from the original on July 28, 2012. Retrieved June 4, 2009.
  125. ^ Dlugokencky, E. J.; et al. (May 2011). "Global atmospheric methane: budget, changes and dangers". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 369 (1943): 2058–2072. Bibcode:2011RSPTA.369.2058D. doi:10.1098/rsta.2010.0341. PMID 21502176. S2CID 8823692.
  126. ^ Saunois, M.; Bousquet, M.; Poulter, B.; et al. (December 12, 2016). "The Global Methane Budget 2000–2012". Earth System Science Data. 8 (2): 697–751. Bibcode:2016ESSD....8..697S. doi:10.5194/essd-8-697-2016. hdl:1721.1/108811. ISSN 1866-3508. Retrieved 28 August 2020.
  127. ^ Saunois, M.; Stavert, A.R.; Poulter, B.; et al. (July 15, 2020). "The Global Methane Budget 2000–2017". Earth System Science Data (ESSD). 12 (3): 1561–1623. Bibcode:2020ESSD...12.1561S. doi:10.5194/essd-12-1561-2020. ISSN 1866-3508. Retrieved 28 August 2020.
  128. ^ Brooks Hays (29 January 2019). "Regulations haven't slowed China's growing methane emissions". UPI. Retrieved 31 January 2019. China's methane emissions increased 50 percent between 2000 and 2015
  129. ^ Miller, Scot M.; Michalak, Anna M.; Detmers, Robert G.; Hasekamp, Otto P.; Bruhwiler, Lori M. P.; Schwietzke, Stefan (January 29, 2019). "China's coal mine methane regulations have not curbed growing emissions". Nature Communications. 10 (1): 303. Bibcode:2019NatCo..10..303M. doi:10.1038/s41467-018-07891-7. PMC 6351523. PMID 30696820.
  130. ^ Guzman, Joseph (2020-03-03). "Exxon calls for tighter regulations of methane". TheHill. Retrieved 2020-03-04.
  131. ^ Alison Durkee (August 10, 2020). "EPA Rescinds Obama-Era Methane Rules As White House Speeds Environmental Rollbacks Ahead Of Election". Forbes.
  132. ^ Emma Newburger (August 29, 2020). "Critics rail against Trump's methane proposal as an 'unconscionable assault on environment'". CNBC.
  133. ^ Jackson, R B; Saunois, M; Bousquet, P; Canadell, J G; Poulter, B; Stavert, A R; Bergamaschi, P; Niwa, Y; Segers, A; Tsuruta, A (14 July 2020). "Increasing anthropogenic methane emissions arise equally from agricultural and fossil fuel sources". Environmental Research Letters. 15 (7): 071002. Bibcode:2020ERL....15g1002J. doi:10.1088/1748-9326/ab9ed2. ISSN 1748-9326.
  134. ^ Niranjan, Ajit (2023-11-15). "EU agrees law to curb methane emissions from fossil fuel industry". The Guardian. ISSN 0261-3077. Retrieved 2024-02-26.
  135. ^ Hutjens, Mike (August 21, 2012). "Use of Rumensin in Dairy Diets". eXtension. Archived from the original on July 9, 2010. Retrieved February 27, 2011.
  136. ^ Bradley, Athena Lee (June 2008). "Manure Management for Small and Hobby Farms" (PDF). Northeast Recycling Council, Inc. Retrieved 2016-12-31.
  137. ^ "Landfill Methane Energy Recovery". Power Partners. December 11, 2009. Archived from the original on September 29, 2015. Retrieved 2016-12-31.
  138. ^ Lizik, William; Im, Jeongdae; Semrau, Jeremy D.; Barcelona, Michael J. (2013). "A field trial of nutrient stimulation of methanotrophs to reduce greenhouse gas emissions from landfill cover soils". Journal of the Air & Waste Management Association. 63 (3): 300–309. doi:10.1080/10962247.2012.755137. PMID 23556240. S2CID 20450110.

External links

This page was last edited on 26 February 2024, at 14:04
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.