Methane - 2022

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

UNIVERSITY OF KARACHI

INSTITUTE OF ENVIRONMENTAL STUDIES

NAME: AYMAN HANIF

FATHER’S NAME: MUHAMMAD HANIF

SEAT NO. H1911002

COURSE TITLE: AIR AND NOISE POLLUTION

COURSE CODE: ENV-502

CLASS: BSc. (Hons), 3RD YEAR

6TH SEMESTER (MORNING PROGRAM)

ASSIGNMENT TITLE: METHANE AND CLIMATE

CHANGE
METHANE AND CLIMATE CHANGE

ABSTRACT:
The most abundant organic trace gas in the atmosphere is methane (CH4). Variations in natural
methane sources were responsible for patterns in atmospheric methane levels recorded in ice
cores in the distant past. Since the 1700s, rapidly increasing human activity, particularly in
agriculture, the use of fossil fuels, and waste management, has more than doubled methane
emissions. Methane is also a greenhouse gas but more potent at trapping radiation than carbon
dioxide. An increase in methane concentration in the atmosphere causes anthropogenic global
warming because the concentration of methane increased due to mainly anthropogenic activities.
Methane is also emitted from natural sources but they account for less emission as compared to
anthropogenic sources. In this review, we look at past patterns in methane concentrations in the
atmosphere, the sources and sinks that influence its growth rate, and the factors that will
influence it in the future. We also discuss the current understanding of methane's effects on
atmospheric chemistry and the relationship between methane and climate change.

1. INTRODUCTION:
Methane (CH4) is a hydrocarbon that is a key component of natural gas. Methane is a colourless,
odourless, and excessively combustible gas. It has a chemical formula CH4 (containing one
carbon atom covalently bonded with four hydrogen atoms). It is also a powerful greenhouse gas
(GHG), hence its presence in the atmosphere has an impact on the global climate. Methane is
emitted from several anthropogenic and natural sources. Anthropogenic sources include
wastewater treatment plants, gas production, thermal power plants, landfills (solid waste disposal
sites), livestock, agricultural activities and some industrial processes. Natural sources include
wetlands, termites, oceanic processes, anaerobic decomposition and geological processes
(volcanoes). (Wuebbles and Katharine, 2002)
It is the third most abundant greenhouse gas after water vapours and carbon dioxide, accounting
for around 20% of global emissions. According to an ice core study, atmospheric CH4
concentrations varied very little over the last 160,000 years, remaining near 0.8 ppm until a
dramatic surge in the last two centuries. Since the beginning of the industrial era, CH4
concentrations have more than doubled, with an average of 1.895 ppm in 2021 (Cheryl Hogue,
2022). Methane was previously thought to have little impact on climate change due to its low
concentration (1.89 ppm) in comparison to carbon dioxide, which has a concentration of 412
ppm (Wuebbles et al. 2017). However, according to a recent study, a single CO2 molecule is
significantly less effective at absorbing terrestrial radiation than a CH4 molecule. Methane is
more than 25 times more effective than carbon dioxide at trapping heat in the atmosphere.
Methane is naturally present in trace levels in the atmosphere, but its concentration has increased
dramatically since the industrial revolution. This rise is significantly faster than the rise in
average CO2.
Changes in atmospheric methane abundance have implications for both chemistry and climate
because methane is both a potent greenhouse gas and a precursor to tropospheric ozone. As a
precursor for tropospheric ozone (O3), influences ozone background levels (Fiore et al., 2002).
Controlling methane has been proven to be a win-win situation, improving both climate and air
quality (Shindell et al., 2012). Methane has increased by a factor of 2.5 from a preindustrial level
of 722±25 ppb (Etheridge et al., 1998; Dlugokencky et al., 2005) to a value of 185±71 ppb in
2017 (Dlugokencky et al., 2017), primarily due to anthropogenic activity (Dlugokencky et al.,
2011). It is eliminated from the atmosphere primarily by reactions with hydroxyl radicals (OH)
in the troposphere, with minor contributions from reactions with excited atomic oxygen (O(1D))
and atomic chlorine (Cl) in the stratosphere and absorbed by soils (Saunois et al., 2016).In the
presence of sufficiently high levels of nitrogen oxides, the oxidation of CH4 by hydroxyl (OH) in
the troposphere leads to the production of formaldehyde (CH2O), carbon monoxide (CO), and
ozone (O3) (NOx). Methane, like CO, helps limit the amount of OH in the troposphere. Methane
also influences the amounts of water vapour and ozone in the stratosphere, and it is essential in
the conversion of reactive chlorine to less reactive HCl in the stratosphere (Wuebbles, 2002).
The focus of this research is to look at past patterns in methane concentrations, the sources and
sinks that influence its growth rate, and the factors that could influence its future growth rate.
This research also looks at how we now understand the consequences of methane on atmospheric
chemistry and climate.

2. RECORD AND BUDGET OF ATMOSPHERIC METHANE CH4:

2.1: ATMOSPHERIC METHANE CH4 IN THE DISTANT PAST:


Air bubbles trapped in glacial regions keep a continuous record of methane concentrations in the
atmosphere. (Blunier et al., 1995, 1998; Brook et al.,1996;Legrand et al., 1988; Chappellaz et al.,
1990; Etheridge et al., 1992, 1998; Jouzel et al., 1993; Nakazawa et al., 1993; Raynaud et al.,
1993).
Ice cores from Greenland and Antarctica have now been used to extend the record of
atmospheric CH4, CO2, and temperature back to 420,000 years ago (Petit et al., 1999), spanning
four glacial-interglacial cycles. Petrenko et al., 2017 discovered that emissions were less than
15.4 Tg CH4 yr1 based on ice with an air age of 11.6 kyr ago.
The record illustrated in Fig. 1 shows two significant aspects of historical CH4 concentrations
that have consequences for future climate. First, current CH4 levels appear to be unparalleled in
history. Ice core data demonstrate glacial-interglacial transitions ranging from 320-350 ppbv to a
maximum of 650-780 ppbv (Petit et al., 1999), corresponding to source increases of up to 50 - 60
TgCH4/year (Brook et al., 1996; Chappellaz et al., 1993a). This is less than half of the current
values, which exceed 1700 ppbv. Second, CH4 levels are strongly connected to atmospheric
temperature records, dropping and rising in lockstep with the temperature at the start and end of
glacial cycles (Petit et al., 1999; Chappellaz et al., 1993a,b; Raynaud et al., 1988).
Fig . 1. Vostoc ice core records of atmospheric temperature and methane concentration from
420,000 years BP to 2000 (petit et al., 1999).

2.2: PRE-INDUSTRIAL TO PRESENT-DAY ATMOSPHERIC METHANE


CONCENTRATION:
Measurements of global CH4 concentrations from polar ice cores during the last 1000 years
indicate numerous interesting patterns. These include (1) fluctuations of 10- 15 ppbv around a
pre-industrial level of 700 ppb; (2) a pole-to-pole gradient of 30- 60 ppb increasing to
approximately 150 ppb in this century, showing continuously higher emissions in the Northern
Hemisphere but variable tropical emissions; (3) a sustained association with temperature
throughout the Little Ice Age around the middle of the millennium; and (4) a strong and
monotonic increase commencing between 1750 and 1800. (Etheridge et al., 1998; Chappellaz et
al., 1997; Blunier et al., 1993; Nakazawa et al., 1993; Khalil and Rasmussen, 1987; Craig and
Chou, 1982). This increase is supported by glacial ice cores dating back to the early 1800s ( Dibb
et al., 1993) and solar spectra dating back to 1950 (Rinsland et al., 1985). According to ice core
data from ground-based stations, the worldwide level of methane has more than doubled since
preindustrial times and is still rising (Rasmussen and Khalil, 1981; Blake and Rowland, 1988;
Dlugokencky et al., 1994a, 1995, 1998; IPCC, 1995, 1996).
Numerous modelling studies suggest that the hydroxyl radical sink for CH4 has decreased from
pre-industrial to the present day. Modelled OH decreases range between 10 to 30%, according to
the model utilized, with the range influenced by a high uncertainty in prior CO, NOx, and
NMHC emissions (Khalil and Rasmussen, 1985; Thompson, 1992; Crutzen and Bruhl, 1993;
Khalil and Rasmussen, 1994b; Osborn and Wigley, 1994; Crutzen, 1995; Brasseur et al., 1998;
Lelieveld et al., 1998; Wang et al., 1998). This range, together with assessments of previous
emissions, indicates that the majority of the observed increase in methane from pre-industrial
times is due to an increase in anthropogenic emissions (Khalil and Rasmussen, 1985, 1994a;
Lelieveld et al., 1993; Subak, 1994; Stern and Kaufmann, 1996).
According to Khalil and Rasmussen (1987, 1994a), there is a strong correlation between rising
population and agricultural emissions and the observed growth in emissions during the last 200
years (Etheridge et al., 1992, 1998; Dlugokencky et al., 1994a).
Studies of human populations and activities over the last millennium, however, indicate that
noticeable anthropogenic effects on CH4 may go back much longer than the start of the pre-
industrial age (Subak, 1994; Blunier et al., 1993). Throughout the late 1970s, methane increased
at an average rate of 20 ppbv/year (Blake and Rowland, 1988). The pace of methane rise did,
however, gradually slow down over the 1980s and 1990s, by an average of 1 ppb/year/year. With
an average growth rate of 8.9 ppbv/year from 1984 to 1996, continuous monitoring of methane
trends in ambient air from 1979 to 1989 revealed a decreasing trend from 20 to 10 ppbv
(Dlugokencky et al., 1994a, 1998; Khalil and Rasmussen, 1993, 1994b; Steele et al., 1992).
However, the amplitude of the seasonal cycle has not shown any significant global trends over
this time (Dlugokencky et al., 1998; Steele et al., 1992), with typical amplitudes around 30 ppb
in the high Southern latitudes, 60 ppb in the high Northern latitudes, and only a small change in
the tropics, according to an analysis of the data for regional and seasonal effects (Khalil et al.,
1993a; Dlugokencky et al., 1994a, 1997).
Following the 1992–1993 anomaly, methane production decreased at a rate of 3 ppbv/year in
1996, when the interhemispheric gradient was 140 ppbv and the world average atmospheric
concentration was 1730 ppbv. Zero growth was anticipated to happen by 2006 when
concentrations gradually levelled at 1800 ppbv if the current downward trend had persisted
(Dlugokencky et al., 1994a; Dlugokencky et al., 1998; Etheridge et al., 1998).
Over the past few decades, the rate of CH4 rise has slowed down over time, which implies that
either methane removal has increased or methane emissions have decreased. Methylchloroform
studies of the worldwide OH concentrations and observed patterns in the seasonal cycle of
methane do not indicate a major change in OH over the past few decades (Dlugokencky et al.,
1997; Prinn et al., 1995). In a similar vein, modelling studies looking into the effects of observed
stratospheric ozone depletion and temperature change over the past ten years discovered that,
while the increase in OH, as a result, is compatible with the observed decrease in trend, these two
factors alone cannot fully explain the decline in growth rate (Bekki and Law, 1997; Fuglestvedt
et al., 1994).
The rising rate of d13C throughout that period is incongruous with a significant increase in the
CH4 sink, according to isotopic analyses (Etheridge et al., 1998). The main cause of the
observed deceleration must therefore be due to diminishing variations in emissions rather than
growing sink strengths. Since there is no known mechanism to account for a long-term decline in
natural resources like wetlands, anthropogenic emissions seem to be the only viable explanation.
To demonstrate that the increase in agricultural sources may be slowing, Khalil and Rasmussen
(1994a,b) look at the shifting link between population and agricultural sources, which are
assumed to be key contributors to CH4 increases from pre-industrial to the present. According to
Steele et al. (1992), the observed abrupt slowing of the growth rate in the Northern compared to
the Southern Hemisphere may be due to human sources, particularly those sources that can be
eliminated swiftly. As long-term CH4 net emissions and OH levels were nearly constant over the
previous decade, Dlugokencky et al. (1998) hypothesized that what had been observed was an
approach to a steady state, even though the precise cause of this long-term decline in the global
methane rate is still unknown.(Wuebbles and Katharine, 2002)
Regardless of the real reason behind this rise, it is important to note at this point how little we
know about the variables that affect methane sources and sinks and how much uncertainty this
adds to projections of future concentrations.
Following the explosion of Mt. Pinatubo in June of 1992 year, methane's growth rate spiked over
six to twelve months. The growth rate dramatically decreased from 1992 to 1993, reaching very
low levels and even zero in some places. Lower stratospheric CH4 concentrations, which lag the
troposphere by four years and had an average growth rate of only 1.8 ppb/year in the Northern
Hemisphere and 7.7 ppb/year in the Southern Hemisphere in 1992, respectively (Dlugokencky et
al., 1994a,b; 1998), resulting in a decrease in atmospheric concentrations (Randel et al., 1999).
Global methane growth rates peaked in 1994 at 8 ppbv/year, then began to decline over the
following years as previously noted (Dlugokencky et al., 1998). Theoretical explanations for this
short-term anomaly include decreasing CH4 clearance rates, reductions in emissions from
anthropogenic or natural sources, and other factors.
Aerosols produced by the Pinatubo eruption prevented UV radiation from reaching the
troposphere, which could have caused a reduction in OH levels and an immediate spike in the
growth rate of methane. According to Dlugokencky et al. (1996), the first, brief increase in
growth rate is compatible with this notion; however, the reasons behind the subsequent fall in
growth rate are yet unknown.
Increases in tropospheric water vapour, an important source of OH, or aerosol-induced
stratospheric ozone depletion, which would increase tropospheric UV radiation and subsequently
lead to an increase in OH formation, have also been proposed as reasons for the short-term drop
(Bekki et al., 1994; Khalil and Rasmussen, 1993).
It has also been suggested that increasing stratospheric temperatures following the Pinatubo
eruption may lead to enhanced mixing of stratospheric air, which has lower CH4 levels
(Schauffler and Daniel, 1994).

3. GLOBAL METHANE BUDGET:


Increased greenhouse gas concentrations have already raised global average surface temperatures
by 1.1 °C above pre-industrial levels, making it difficult to stabilize the climate (World
Meteorological Organization 2019). Atmospheric methane concentrations on average across the
globe increased to 1875 parts per billion (ppb) at the end of 2019—more than 2.5 times pre-
industrial levels (Dlugokencky 2020). Methane has a significant global warming potential that
seems to be between 28 and 36 times more than that of Carbon dioxide over 100 years (IPCC,
2021). Despite having a relatively short perturbation lifetime (12.4 years), methane is far less
prevalent in the atmosphere than CO2 (Balcombe et al., 2018).
Since 1750, emissions of nitrous oxide (N2O), carbon dioxide (CO2), and methane (CH4) have
contributed almost one-fourth of the total radiative forcing.(Etminan et al 2016). For the 2008–
2017 decade, the ensemble of 22 inversions inferred total emissions of 576 Tg CH4 yr1, with the
highest ensemble mean emission of 596 Tg CH4 yr1 [572–614] for 2017 (Marielle Saunois et al.,
2020).
It is critical to comprehend and quantify the global methane budget to evaluate workable climate
change mitigation strategies. As a result of rising atmospheric emissions and concentrations,
CH4 has overtaken carbon dioxide as the second-most significant human-influenced greenhouse
gas in terms of climate forcing.
Pre-industrial emissions were roughly 220 Tg CH4 yr1 based on observations of CH4 abundance
from ice cores and firn and assuming its lifetime was equivalent to that of today (9 years).
Wetland emissions were the main source back then, but estimates of anthropogenic fossil
emissions in the current budget are affected by a great deal of uncertainty on emissions from
natural fossil sources. An isotope mass balance can be used to determine how much of the total
CH4 emissions come from fossil sources. That isotope is δ13CCH4 (Schwietzke et al., 2016)
Measuring the global distribution of surface methane since 1983 has shown that atmospheric
methane nearly reached a stable state between 1983 and 2006 and has since resumed its upward
trend. According to Nisbet et al. (2014) and Dlugokencky et al. (2018), methane growth rates fell
from 12 ppb yr1 during 1984–1991 to 5 ppb yr1 during 1992–1998 and to 0.7–0.6 ppb yr1
during 1999–2006. (Dlugokencky et al., 2018). After 2006, methane levels began to rise once
more, growing at a pace of 5.7 ppb per year from 2007 to 2013, reaching 12.6 ppb per year in
2014 and 10.0 ppb per year in 2015. (Nisbet et al., 2016; Dlugokencky et al., 2018).
Although anthropogenic activities are widely believed to be to blame for the long-term rise in
methane since the dawn of civilization (Dlugokencky et al., 2011), there is disagreement
regarding the causes of the methane stabilization between 1999 and 2006 and the subsequent
resurgence of growth since 2007. The stabilization between 1999 and 2006 has been attributed in
previous studies to the interaction of rising anthropogenic emissions with falling wetland
emissions (Bousquet et al., 2006), declining fossil fuel emissions (Dlugokencky et al., 2003;
Simpson et al., 2012; Schaefer et al., 2016), stabilizing emissions from microbial and fossil fuel
sources (Levin et al., 2012), or shifting methane sinks (Rigby et al., 2008; Montzka et al., 2011;
Schaefer et al., 2016). Alternative explanations for the observed renewed growth since 2007
include increases in tropical emissions, such as agricultural emissions (Houweling et al., 2014;
Nisbet et al., 2016), tropical wetland emission levels (Bousquet et al., 2011; Maasakkers et al.,
2019), rises in emission from fossil fuels (Rice et al., 2016; Worden et al., 2017), and decreases
in sources offset by decreases in (Saunois et al., 2017).
During the 2008-2017 decade, five regions—China, Southeast Asia, the United States, South
Asia, and Brazil—accounted for more than 40% of total CH4 emissions.

4. TOTAL METHANE BUDGET OF PAKISTAN:


Pakistan is an agricultural country. Agriculture and livestock generate 21% of the GDP, employ
45% of the labour force, and contribute to the country's foreign reserves. In recent decades,
livestock has surpassed crops in agricultural value addition. (Abas et al., 2017)
GHG emissions can be ascribed to population growth rates, increased living standards
demonstrated by higher per capita income, and net export earnings (Štreimikienė & Balezentis,
2016). Because most developing countries rely primarily on agriculture, lower cropland
productivity due to higher CO2 concentrations will have a detrimental impact on socioeconomic
growth. Tropical countries will be struck the worst by climate change, which will reduce the
productivity of their lands, the majority of which are arid ( Khan and Abas, 2012). Pakistan has a
large area of barren territory that has been rendered cultivable by significant irrigation over the
Indus basin during the last century. The emissions from agriculture are caused by flooding rice
fields.
The emissions from agriculture are caused by flooding rice fields. Domestic cattle contribute to
emissions owing to enteric fermentation, which happens naturally in regularly bred cattle such as
goats, cows, and sheep. Different animals emit different amounts of methane depending on their
food intake and digestive processes, though the key influences are the animal's age and the
amount of feed it consumes. Along with cattle, flooded rice fields are a major source of
emissions due to the anaerobic breakdown of organic waste in them, where the gas escapes via
diffusion. Pakistan, India, and Bangladesh are important rice-growing countries, and their rice
fields account for a large portion of their emissions. Livestock is responsible for around 67% of
agricultural emissions, while manure management is responsible for 6%. Methane is the most
significant component of these emissions (Abas et al., 2017).
According to certain predictions, Pakistan's population would reach 350 million by 2050. The
agricultural and livestock markets are predicted to expand significantly in light of the economic
growth expected from increasing commerce, particularly under the auspices of CPEC and the
One Belt One Road plan (Abas et al., 2017)
GHG reductions in Pakistan may include improving cattle feed, agricultural management,
reducing methane emissions from rice agriculture, and boosting productivity and efficiency. The
Alternative Energy Development Board (AEDB), the National Energy Conservation Center
(Enercon), and the Ministry of Climate Change are collaborating with academia to improve
energy conservation and renewable energy. A multi-nutrient feed block can lower methane
emissions by up to 23% per animal. High protein feed can enhance cattle emissions by 40% at
the expense of an additional 85% energy in cattle raising. Better soil, water, and fertiliser
management can help to reduce nitrous oxide emissions (Tahir et al., 2015)
Pakistan must enhance its irrigation practices, fertiliser and manure management, plant residue
and solid waste management to increase efficiency and minimise emissions. Rice varieties can be
genetically modified to reduce methane emissions from paddy crops. Furthermore, burning
leftover wheat and rice straw, as well as cotton sticks, raises overall GHG emissions
(Koneswaran and Nierenberg, 2008)

5. SOURCES AND SINKS:

5.1: SOURCES:
In contrast to carbon dioxide, methane is emitted into the atmosphere from a range of natural and
man-made sources. Biogenic sources connected to agriculture and waste management, such as
anaerobic fermentation, both animal and human waste, rice paddies, biomass burning, and
landfills, produce anthropogenic emissions. Exploiting fossil fuels like natural gas, coal, and
petroleum also releases methane into the atmosphere. Wetlands, termites, other wild ruminants,
oceans, and hydrates produce natural methane emissions.
Because there are so many different sources of methane, a wide range of variables, such as
energy use, population distribution, agricultural methods, and climate, have an impact on
emissions.(Wuebbles and Katharine, 2002)
Major sources of methane emission are wetlands, biomass burning, combustion of fossil fuel and
agricultural activities.(Schwietzke et al., 2016)

5.1.1: NATURAL SOURCES:


Natural sources account for 40% of the total methane emission(Schwietzke et al., 2016).
Methane emissions from natural sources are mainly caused by anaerobic decomposition by
bacteria in flooded fields, landfills or other dumpsites, domestic ruminant digestive tracts like
those of cattle or sheep, wild ruminant digestive tracts like those of buffalo or termites, and even
human digestive tracts. Methane production is temperature-dependent, with a maximum growth
temperature between 37 and 45 degrees C. (Boone, 2000). Today there is more production of
methane due to global temperature rise. That means as the temperature increases, more
production of methane naturally, increases emission by these sources and finally enchanting
warming of the earth as the presence of methane will increase in the atmosphere.
Other minor sources of CH4 production include photochemical processes and aerobic bacterial
methylation of substances (Florez-Leiva et al., 2013).
Brazil, Canada, Equatorial Africa, Russia, and Southeast Asia dominate natural emissions.
(Stavert et al., 2022).
Following are the natural sources of methane:

A. WETLANDS:
Wetlands, including bogs, marshes, fens, and permafrost, offer an environment that is suitable
for the bacteria that create methane as organic matter decomposes.
A wide range of environmental factors affects the emissions from natural wetlands. These
characteristics include plenty of nutrients, main carbon in soil(Cao et al., 1996b; Miller et al.,
1999; Smith et al., 2000; Yavitt et al., 2000), vegetation type and cover (King et al., 1998;
Bellisario et al., 1999; Joabsson et al., 1999; Van der Nat and Middelburg, 2000), anaerobic
condition, and last one and the most important one which provides an anaerobic condition that is
water table depth and soil temperature.(Wuebbles and Katharine, 2002)

B. TERMITES:
According to estimates, 13% of the world's methane emissions from natural sources come from
termites and their mounds. As part of their regular digestion, termite gut microbes produce
methane, albeit the amount produced varies depending on the species. The termite population,
which varies greatly across different geographic areas, has a considerable impact on
emissions.(Wuebbles and Katharine, 2002)

C. OCEANS:
Approximately 6% of the world's natural methane emissions are thought to originate from the
oceans. The anaerobic digestion of fish and zooplankton in the water is one known source of
methane from the oceans, coupled with methane production in sediments and drainage areas near
coastal locations.

D. GEOLOGICAL PROCESSES:
The amount of methane released from natural geological sources is seeps and mud volcanoes.
Several gasses are released into the atmosphere by geological sources, including SO2, H2O
(water vapour), HCl, and CO2 from volcanoes and H2, NO2, CO2, and CO2 from hydrothermal
vents (Judd et al., 2002). Methane is one of these gasses which emit during volcano eruptions,
from hydrothermal vents and fissures.

E. METHANE HYDRATES:
Methane hydrates are particularly important among the vast reserves of CH4 carbon that
naturally influence the ocean-atmosphere system and subsequently, the global climate. Methane
hydrate comprises only a small percentage of atmospheric methane but may be of great
importance in future (Ruppel et al., 2017).
When water and low-molecular-weight gasses (CO2, H2S, CH4, and higher-order hydrocarbons)
mix in a clathrate structure, a substance resembling ice is created that is known as a gas hydrate
(Sloan and Koh, 2008). Methane predominates over all other gasses in natural gas hydrates that
have been directly tested, which may partly be due to the abundance of CH4 there rather than
higher-order thermogenic gasses at the shallow depths that coring and drilling normally access.
At standard temperature and pressure, 1 m3 of gas hydrate can sequester up to 180 m3 of
methane. Methane hydrate concentrates CH4 inside its cage-like molecules (STP) (Ruppel et al.,
2017).
There are large methane hydrate concentrations on arctic islands, and these deposits have been
investigated as possible fuel sources. Since methane is a greenhouse gas, there has been fear that
any increase in temperature could cause some of the hydrate to melt and release significant
volumes of the gas into the atmosphere, resulting in the positive effect of promoting further
warming (Cao, 2002).

Fig. 2. Methane hydrate (white material) on the floor of Northern Gulf of Mexico.
The photograph was taken in 2014 by the Deep Discoverer remotely-operated vehicle managed
by the National Oceanic and Atmospheric Administration’s Ocean Exploration and Research
Program. Credit: NOAA

5.1.2: ANTHROPOGENIC SOURCES:


Anthropogenic sources accounts for about 60% of the total global methane emission.
Anthropogenic sources are discussed below:

A. RICE CULTIVATION:
conditions and farming methods used during cultivation, including weather, soil properties, and
farming methods such as water management, fertilizers, and other additives, as well as varieties
of plants. (Minami and Neue, 1994; Cao et al., 1995, 1996a; Huang et al., 1997, 1998; Neue et
al., 1997; Sass and Fisher, 1997; Khalil et al., 1998; Neue and Roger, 2000).
Rice fields provide ideal conditions for the methanogens for the production of methane. It has
been discovered that techniques like infrequent field draining, the addition of oxidants or other
mineral fertilizers, and the choice of low CH4 cultivars each reduce emissions by roughly 40–
55%, 20–70%, and sometimes even up to 60%, respectively. (Butterbachbahl et al., 1997; Cole et
al., 1997; Minami, 1997; Sass and Fisher, 1997; Mitra et al., 1999).Alternately, adding organic
fertilizers could increase emissions by more than 50% when compared to non-organic fertilizers
(Buendia et al., 1997; Denier van der Gon and Neue, 1995; Yagi et al., 1997).

B. LIVESTOCK:
Cattle and other ruminants including other domesticated animals, such as goats, sheep, and
buffalo emit methane and this methane is directly related to their diet (Cole et al., 1997; Harper
et al., 1999). Methane results as a byproduct of incomplete digestion. (Wuebbles and Hayhoe,
2002). Generally, higher quality diets will enable animals to more fully digest their food,
enhancing protein absorption while lowering CH4 emissions. This is especially true for ruminant
diets in impoverished nations, where raising the comparatively low quality of bovine feed could
result in up to 75% lower emissions per kilogramme of milk produced (Ward et al., 1993;
Crutzen et al., 1986).

C. WASTEWATER TREATMENT:
Soluble organic matter, suspended particles, pathogenic organisms, and chemical pollutants are
all taken out of sewage wastewater during treatment. If organic components in the wastewater
are processed without oxygen and the resulting methane is discharged into the atmosphere, these
treatment procedures may result in methane emissions.

D. LANDFILLS:
As garbage decomposes anaerobically in landfills and open dumps, methane is produced
(oxygen-free). The volume and moisture level of the waste, as well as the site's architecture and
management procedures, all affect how much methane is produced. Dry areas have landfills that
are less productive than ones with high moisture content. Several environmental and
technological factors, such as the temperature, moisture levels, and CH4 concentration inside the
landfill, the volume, organic composition, and age of the garbage, and the thickness of the
covering layer, all have an impact on landfill emissions (Peer et al., 1993; Czepiel et al., 1996;
Bogner and Spokas, 1993).

E. FOSSIL FUEL COMBUSTION:


The anaerobic decomposition of buried dead creatures is one of the natural processes that
produce fossil fuels. Natural gas, oil, and coal are examples of fossil fuels with high carbon
content.
F. BIOMASS BURNING:
Numerous contaminants are released into the atmosphere when biomass is burned. When
combustion is complete, carbon dioxide makes up the majority of the emissions. However,
smouldering fires and incomplete combustion can result in significant volumes of CH4 and other
higher-order hydrocarbons being created. Methane emissions from burning biomass are
influenced by the stage of combustion attained, the biomass's carbon content, and the volume of
the biomass burned.

G. GAS PRODUCTION:
Natural gas processing, transmission, and distribution leaks are the primary cause of methane
emissions linked to fossil fuels because natural gas contains over 90% methane ( Beck et al.,
1993; EPA, 1993a).

H. COAL PRODUCTION:
When gas that was trapped between layers of coal during its formation is freed as the coal is
mined, methane is also produced from coal mines. Estimates of CH4 emissions from coal and
mines at the global and even regional levels depend on many assumptions about the type of coal,
the depth of the mine, mining techniques, the amount of methane in the coal seam, and whether
methane is flared or released.(Beck et al., 1993; Kirchgessner, 2000)
Methane emissions from mining activities and natural gas consumption may be more easily
controlled by changing mining procedures and enhancing gas leakage controls than by
abandoning coal or gas use in the absence of suitable alternative energy sources. Methane
released from these sources might potentially be captured and used as an energy source, which
would have further advantages.

5.2: SINKS:
Only one main and two minor sinks for tropospheric methane exist in contrast to the numerous
sources of gas.
The hydroxyl radical oxidation of methane in the troposphere is the primary method by which it
is removed from the earth's atmosphere (OH). A negatively charged oxygen atom joined to a
hydrogen atom forms the hydroxyl radical (OH). Hydroxyl radicals are a form of "sink" because
they "sweep" the atmosphere clean of damaging substances and destroy them. This is why OH is
dubbed the "atmosphere cleanser.". This is why OH is referred to as the "atmosphere cleaner."
Atmospheric methane undergoes a protracted chain of chemical processes before interacting with
OH to produce CO2.) Some of the methane in the troposphere diffuses into the stratosphere,
where the exact mechanism purges the atmosphere of methane.
The concentration of hydroxyl radical (OH) is the most significant factor in the rate at which
methane is removed from the atmosphere since it reacts with OH to remove around 500
TgCH4/year (nearly 90% of the total sink). The remaining CH4 is eliminated through dry soil
oxidation, which removes 30 TgCH4/year about 5%, and transfers to the stratosphere, which was
previously estimated to remove 40 TgCH4/year about 7%based on data from the Upper
Atmosphere Research Satellite to be only 27 TgCH4/year.
(Gettelman et al., 1997).
Ozone and water vapour in the troposphere photodissociate to produce OH. Most tropospheric
contaminants, including carbon monoxide, NOx species, and organic molecules, are oxidized
primarily by it (Crutzen, 1995). Methane, which is the most prevalent organic species in the
atmosphere, has a significant impact on defining the tropospheric oxidizing capacity since it
starts a significant number of chemical processes. Nearly 90% of methane degradation now takes
place in the troposphere as a result of the oxidation of methane by OH, which is a very efficient
process (IPCC, 1996). OH, levels and the reaction's rate constant limit how much methane can
be extracted.
Methane emissions directly and its oxidation products, especially carbon monoxide, have an
impact on OH concentrations. Based on CH3CCl3 analyses and a thorough understanding of the
loss rate with OH, methane has an atmospheric lifetime of 8.9F0.6 years. (Prinn et al., 1995)
Methane is eliminated by OH on an "adjustment" time scale that surpasses the atmospheric
lifetime by about 3 years, even though most trace gasses have a turnover time that explains how
long they are present in the atmosphere.
In addition to interactions with CH4, the hydroxyl radical is also eliminated by interactions with
its byproduct, CO. In a chemical feedback cycle involving CH4, CO, and OH, even a slight
increase in background methane due to increasing emissions will cause OH to decrease and CO
to be produced. A further decrease in OH levels will result from OH's subsequent oxidation of
the carbon monoxide that was created by the CH4 oxidation. The lifespan of methane increases
as a result of the additional drop in OH, amplifying the initial disruption.
As a result, there can be a positive feedback loop whereby rising methane emissions can cause
the troposphere's general oxidizing ability to decline, methane removal to slow down, and
methane concentrations to subsequently rise.
Even though there may be no apparent quantifiable change in OH, an incremental perturbation in
CH4 will return to an equilibrium concentration at a later rate than anticipated from the
atmospheric lifetime. Naturally, methane emission from sources is balanced by its removal by
the sink. But due to human activities, the concentration of methane increases rapidly. This causes
anthropogenic global warming which makes favourable conditions for more methane production.
Even though it is much less than the OH sink, it is also anticipated that the soil sink for methane
may vary in the future. Temperature and soil water content are two factors that affect methane
oxidation in soil (Mancinelli, 1995; King, 1997), and climate change may have an impact on
both of them. Perhaps more significantly, compared to untreated soils, arable land absorbs CH4
at a far lower rate, especially when fertilised (Boeckx et al., 1997; Mosier et al., 1997b, Powlson
et al., 1997). The soil sink has already been reduced by changes in land use, such as the
conversion of grasslands and forests to farmland, and this trend is likely to continue in the future
(King, 1997; Mosier et al., 1997a). As more grasslands and forests are turned over to agriculture,
the soil sink has already declined and is projected to do so in the future (King, 1997; Mosier et
al., 1997a). Previously farmed land continues to have a lower rate of oxidation than unspoiled
areas even after being returned to its natural state (e.g., Dobbie and Smith, 1996; Hudgens and
Yavitt, 1997; Prieme et al., 1997). Given that expanding populations continue to put a strain on
natural ecosystems, the apparent immutability of human impacts on CH4 oxidation rates has
significant ramifications for the future of land management techniques.

6. IMPACT OF CH4 ON TROPOSPHERIC AND STRATOSPHERIC


CHEMISTRY:

6.1: METHANE AND TROPOSPHERE:


As previously discussed, methane oxidation is one of the key reaction pathways influencing
hydroxyl concentrations in the atmosphere on a global scale. Methane oxidation can be either a
production or destruction process for odd hydrogen (OH+HO2), depending on nitric oxide levels.
Thus, various chemically coherent zones (Thompson et al., 1989) can be differentiated based on
nitrogen oxide concentrations.
The methane oxidation cycle is a significant carbon monoxide source, accounting for nearly one-
quarter of the carbon monoxide in the troposphere. Carbon monoxide concentrations are much
more variable than methane concentrations due to their relatively short atmospheric lifetime
(about 1- 3 months) and the range of natural and manmade sources that contribute to their
budget. These sources include the combustion of fossil fuels, biomass burning, and the oxidation
of natural hydrocarbons released by vegetation (e.g., isoprene). The carbon monoxide oxidation
cycle, like the methane oxidation cycle, is affected by the amount of nitric oxide in the
environment.
Formaldehyde is a significant step in methane and other hydrocarbon elimination processes, as
well as in the overall chemical reactivity of the troposphere. Nonmethane hydrocarbons and
higher aldehydes are the other major photochemical precursors of formaldehyde. Non-methane
hydrocarbons are emitted from both natural and anthropogenic sources, whereas higher
aldehydes are created in situ by photochemistry. There are also significant anthropogenic sources
of formaldehyde (e.g., automobile exhaust). The troposphere contains around 10% of the ozone
in the atmosphere. Historically, the principal source of tropospheric ozone was assumed to be the
downward movement of ozone from the stratosphere (Crutzen, 1988). The net tropospheric
photochemical synthesis of ozone is now widely acknowledged to be of comparable magnitude
to the downward transport source ( Fishman, 1985; WMO, 1985, 1995; Isaksen, 1988; Penkett,
1988).
Nitrogen dioxide photolysis is by far the most important photochemical pathway for creating
ozone in the troposphere. This suggests that the rate of creation is roughly proportional to the
concentration of nitric oxide. High NOx concentrations over the continental boundary layer
indicate that this location is most likely a net generator of ozone. Increased NOx emissions may
result in increased ozone levels, particularly in the tropics.
6.2: METHANE AND STRATOSPHERE:
Although 90% of total methane emissions are consumed by interaction with tropospheric
hydroxyl and around 5% by soil uptake (IPCC, 1995), the remaining 5% of methane flux enters
the stratosphere (Gettelman et al., 1997).
The interaction with OH remains the dominating sink for CH4 in the stratosphere and higher
layers of the atmosphere, but reactions with chlorine atoms and excited oxygen atoms are also
important. Chlorine atom reactions account for approximately 9% of methane loss in these places
(Brenninkmeijer et al., 1995). Measurements also indicate that greater methane oxidation caused
by increasing chlorine concentrations in the stratosphere has resulted in increased formation of
stratospheric OH (e.g., Burnett and Burnett, 1995).
The stratosphere contains roughly 90% of the ozone in the atmosphere. Ozone formation in the
stratosphere begins with the photodissociation of oxygen (O2). This process yields two ground-
state oxygen atoms that can react with oxygen to form ozone. The catalytic processes of chlorine
and bromine are very effective at destroying ozone.
Methane is significant in stratospheric chlorine chemistry, acting as both a source and a sink in
key processes involving reactive chlorine. The principal source of hydrochloric acid, the main
chlorine reservoir species, is the direct interaction of methane with a chlorine atom.

7. IMPACT OF CH4 ON OZONE:


The distribution of ozone in the troposphere and stratosphere is currently being impacted by
changing atmospheric quantities of several substances, including methane. The combined
impacts of rising methane and other gas concentrations on ozone have been studied in several
research investigations using numerical models. Stordal and Isaksen (1987); Wuebbles et al.
(1983). Increasing methane concentrations, according to numerical models of atmospheric
chemical and physical processes, result in net ozone synthesis in the troposphere and lower
stratosphere, and net ozone destruction in the high stratosphere.(Owens et al., 1982, 1985;
WMO, 1985, 1991, 1995; Isaksen and Stordal, 1986). The total result of these computations is
that methane causes a net increase in ozone. Published effects on the predicted change in total
ozone range from +0.3% (Prather in WMO, 1985) to +4.3% for a doubling of methane
concentration (early articles evaluated CH4 rises from 1.6 to 3.2 ppmv, whereas subsequent
assessments assume a globally-averaged change from 1.7 to 3.4 ppmv) (Owens et al., 1985). The
published model findings tend to be at the upper end of this range when radiative feedback
effects are incorporated (accounting for temperature variations in the stratosphere) (Owens et al.,
1985; WMO, 1985; Isaksen and Stordal, 1986).

8. METHANE AND CLIMATE CHANGE:


In the past, concern about human activities affecting global climate has mostly focused on
carbon dioxide because of its importance as a greenhouse gas and the quick rate at which its
atmospheric concentration has been rising. However, other greenhouse gases have had a
considerable impact on climate over the last few centuries. Rising methane concentrations,
second only to carbon dioxide, have contributed to a significant increase in radiative forcing
since pre-industrial times. The combined effect of methane and other greenhouse gases has
almost doubled the overall increase in greenhouse radiative forcing on climate compared to CO2.
Methane, like other greenhouse gases, absorbs infrared radiation emitted by the relatively warm
surface and emits it to space at colder atmospheric temperatures, resulting in a net trapping of
infrared radiation in the atmosphere. This is known as the greenhouse effect. The net radiative
forcing on climate is determined by the balance of absorbed solar energy and emitted infrared
radiation.
Despite having a lower atmospheric abundance than carbon dioxide, methane is a significant
greenhouse gas. Donner and Ramanathan (1980) found that with current levels of methane, the
globally averaged surface temperature is roughly 1.3 K higher than it would be without methane.
On a molar basis, an additional mole of methane in the current atmosphere is approximately 24
times more effective than an additional mole of carbon dioxide at absorbing infrared radiation
and influencing climate (WMO, 1999).
The strongest bands for methane absorption in the infrared are towards the 'window region"s
short wavelength edge (∼7 –13 Am, where there is little absorption by H2O or CO2). The 7.66
Am (1306 cm1) absorption band is the most prominent infrared spectral characteristic of
methane. Methane radiative forcing increases about as the square root of its concentration due to
methane saturation of the line cores and emissions from the pressure-broadened Lorentz line
wings (IPCC, 1990; Wigley, 1987). The efficacy of methane absorption is additionally affected
by overlapping with absorption by water vapour and other species (especially nitrous oxide).
Wang et al. (1991) demonstrate that the greenhouse forcing of methane has distinct geographical
impacts on climate than carbon dioxide and that methane must be explicitly accounted for when
attempting to model climate.
Increasing methane concentrations are expected to account for a large portion of the rise in
radiative forcing caused by greenhouse gases during the previous two centuries. IPCC (1996)
(1996). Other research investigations have revealed similar percentages for the effect of methane
over this period (Rodhe, 1990; Hansen et al., 1989; MacKay and Khalil, 1991).

9. METHANE MITIGATION:
Methane is emitted by cattle and other ruminants, including domesticated animals such as goats,
sheep, and buffalo, and this methane is directly tied to their food. Methanogens are present in
these ruminants and responsible for methane emissions. But a high-quality diet can reduce
emissions from this source. This can be done by giving a protein-rich diet to animals, which
improves their digestion. Methane mitigation in ruminants can also be achieved by
biotechnology. Several biotechnology solutions are being investigated at the moment. In
Australian sheep, a vaccine against three chosen methanogens reduced CH4 output by about 8%.
(Wright et al., 2004).
Some additives can be included in the diet to manage methane emissions. Either they improve
metabolism or inhibit methanogens. Among feed additives, ionophore antibiotics such as
monensin and lasalocid, which are generally employed to improve animal production efficiency,
have been shown to reduce CH4 generation (reviewed by Beauchemin et al. (2008)).
Methane can also be reduced by proper management of landfills, where disposal of municipal
water takes place. These places are giant factories for methane production. Dumping of waste
regularly provides organic matter for decomposition. In landfills, there is the hype of solid water,
which provide the suitable condition for methane production as they limit airflow which
eventually limits the availability of oxygen for aerobic decomposition. So there is a
decomposition of organic matter but anaerobically due to the absence of oxygen. Alternatively,
methane can be captured and used as a source of energy. To do this there is a need for a setup
which provides a completely closed place and technology. This brings methane as a source of
energy and compost as a byproduct.
Pakistan is an agricultural country. More than 60% of the population depends on it for income.
Rice is a major crop in Pakistan. Rice cultivation emits methane into the atmosphere which
influences global warming. Rice required fields of stagnant water for growth. Rice field with
water provides similar condition as wetlands. Due to climatic conditions and the availability of
nutrients and most importantly the absence of air, there is the production of methane. Methane
production depends on many soil characteristics like water table depth, soil nutrients, fertilizers,
and different varieties of vegetation. Methane emissions can be reduced by using chemical
fertilizers instead of organic fertilizers. It can also be reduced by taking those varieties which
limit methane production.

10.CONCLUSION:
Methane concentrations in the atmosphere have risen considerably during the previous century
and continue to rise. While the budget has proven the causative involvement of human activities
in this increase, there are major difficulties in understanding the factors that govern emissions
from diverse sources and how they will evolve.
The continuation of current emissions commits us to higher future concentrations, and the longer
emissions continue to rise, the larger the reductions required to stabilise concentrations at a
specific stage. As previously stated, higher temperatures are required for methane production.
Today's global temperature is substantially higher than it was a century ago, and this state acts as
an industry, producing more and more methane. That is a positive thing for global warming.
Most anthropogenic activities have resulted in increased methane concentrations today. Methane
emissions can be reduced in a variety of ways, as previously discussed.
Further research on the sources and mitigation strategies for reducing methane emissions is
required.
REFERENCES:
• Abas, N., Kalair, A., Khan, N., & Kalair, A. R. (2017). Review of GHG emissions in
Pakistan compared to SAARC countries. Renewable and Sustainable Energy
Reviews, 80, 990-1016.
• Beck, L.L., Piccot, S.D., Kirchgessner, D.A., 1993. Industrial sources. In: Khalil, M.
(Ed.), Atmospheric Methane: Sources, Sinks and Role in Global Change. Springer-
Verlag, New York, NY, pp. 341 – 399.
• Bekki, S., Law, K., 1997. Sensitivity of the atmospheric CH4 growth rate to global
temperature changes observed from 1980 to 1992. Tellus 49B, 409 – 416.
• Blake, D.R., Rowland, F.S., 1988. Continuing worldwide increase in tropospheric
methane, 1978 to 1987. Science 239, 1129 – 1131.
• Brook, E. J., Sowers, T., & Orchardo, J. (1996). Rapid variations in atmospheric
methane concentration during the past 110,000 years. Science, 273(5278), 1087-1091.
• Blunier, T., Chappellaz, J., Schwander, J., Stauffer, B., & Raynaud, D. (1995). Variations
in atmospheric methane concentration during the Holocene epoch. Nature, 374(6517),
46-49.
• Boone, D., 2000. Biological formation and consumption of methane. In: Khalil, M. (Ed.),
Atmospheric Methane: Its Role in the Global Environment. Springer-Verlag, New York,
NY, pp. 42 – 62.
• Bogner, J., Spokas, K., Burton, E., Sweeney, R., Corona, V., 1995. Landfills as
atmospheric methane sources and sinks. Chemosphere 31, 4119 – 4130.
• Burnett, E.B., Burnett, C.R., 1995. Enhanced production of stratospheric OH from
methane oxidation at elevated reactive chlorine levels in northern midlatitudes. J. Atmos.
Chem. 21, 13 – 41.
• Blunier, T., Chappellaz, J., Schwander, J., Barnola, J., Desperts, T., Stauffer, B.,
Raynaud, D., 1993. Atmospheric methane, record from a Greenland ice core over the last
1000 years. Geophys. Res. Lett. 20, 2219 – 2222.
• Buendia, L., Neue, H., Wassmann, R., Lantin, R., Javellana, A., 1997. Understanding the
nature of methane emission from rice ecosystems as basis of mitigation strategies. Appl.
Energy 56, 433 – 444.
• Brasseur, G., Kiehl, J., Muller, J.-F., Schneider, T., Granier, C., Tie, X., Hauglustaine, D.,
1998. Past and future changes in global tropospheric ozone: impact on radiative forcing.
Geophys. Res. Lett. 25, 3807 – 3810.
• Balcombe, P., Speirs, J. F., Brandon, N. P., & Hawkes, A. D. (2018). Methane emissions:
choosing the right climate metric and time horizon. Environmental Science: Processes &
Impacts, 20(10), 1323-1339.
• Bellisario, L.M., Bubier, J.L., Moore, T.R., Chanton, J.P., 1999. Controls on CH4
emissions from a northern peatland. Global Biogeochem. Cycles 13, 81 – 91.
• Butterbachbahl, K., Papen, H., Rennenberg, H., 1997. Impact of gas transport through
rice cultivars on methane emission from rice paddy fields. Plant Cell Environ. 20, 1175 –
1183.
• Cao, Z. 2002. Modeling of gas hydrates from first principles. Doctoral dissertation,
Massachusetts Institute of Technology.
• Cao, M., Marshall, S., Gregson, K., 1996b. Global carbon exchange and methane
emissions from natural wetlands: application of a process-based model. J. Geophys. Res.
101, 14399 – 14414.
• Cao, M., Dent, J., Heal, O., 1995. Modeling methane emissions from rice paddies. Global
Biogeochem. Cycles 9, 183 – 195.
• Cole, C., Duxbury, J., Freney, J., Heinemeyer, O., Minami, K., Mosier, A., Paustian, K.,
Rosenberg, N., Sampson, N., Sauerbeck, D., Zhao, Q., 1997. Global estimates of
potential mitigation of greenhouse gas emissions by agriculture. Nutr. Cycling
Agroecosyst. 49, 221 – 228.
• Cheryl Hogue. (2022). Atmospheric levels of methane post a record rise in
2021. Chemical & Engineering News, 19–19.
• Chappellaz, J., Barnola, J. M., Raynaud, D., Korotkevich, Y. S., & Lorius, C. (1990). Ice-
core record of atmospheric methane over the past 160,000 years. Nature, 345(6271), 127-
131.
• Chappellaz, J., Blunier, T., Raynaud, D., Barnola, J., Schwander, J., Stauffer, B., 1993a.
Synchronous changes in atmospheric CH4 and Greenland climate between 40-kyr and 8-
kyr BP. Nature 366, 443 – 445.
• Chappellaz, J., Fung, I.Y., Thompson, A.M., 1993b. The atmospheric CH4 increase since
the Last Glacial Maximum. Tellus 45B, 228 – 241.
• Craig, H., Chou, C.C., 1982. Methane: the record in polar ice cores. Geophys. Res. Lett.
9, 1221 – 1224.
• Crutzen, P., Bruhl, C., 1993. A model study of the atmospheric temperatures and
concentrations of ozone, hydroxyl, and some other photochemically active gases during
the glacial, the preindustrial Holocene and the present. Geophys. Res. Lett. 20, 1047 –
1050.
• Crutzen, P., 1995. Overview of tropospheric chemistry: developments during the past
quarter century and a look ahead. Faraday Discuss. 100, 1 – 21.
• Crutzen, P., Aselmann, I., Seiler, W., 1986. Methane production by domestic animals,
wild ruminants, and other herbivorous fauna and humans. Tellus 38B, 271.
• Czepiel, P., Mosher, B., Crill, P., Harriss, R., 1996. Quantifying the effect of oxidation on
landfill methane emissions. J. Geophys. Res. 101, 16721 – 16729.
• Dibb, J., Rasmussen, R., Mayewski, P., Holdsworth, G., 1993. Northern hemisphere
concentrations of methane and nitrous oxide since 1800: results from the Mt. Logan and
20D ice cores. Chemosphere 27, 2413 – 2423.
• Dlugokencky, E. J. (2005). Conversion of NOAA atmospheric dry air CH4mole fractions
to a gravimetrically prepared standard scale. Journal of Geophysical Research, 110(D18).
• Dlugokencky, E. J., Nisbet, E. G., Fisher, R., & Lowry, D. (2011). Global atmospheric
methane: budget, changes and dangers. Philosophical Transactions of the Royal Society
A: Mathematical, Physical and Engineering Sciences, 369(1943), 2058–2072.
• Dlugokencky, E. J., Lang, P. M., Crotwell, A. M., Mund, J. W., Crotwell, M. J., &
Thoning, K. W. (2017). Atmospheric methane dry air mole fractions from the NOAA
ESRL carbon cycle cooperative global air sampling network, 1983-2016. Division N-
EGM, Editor. Version, 07-28.
• Dlugokencky, E., Steele, L., Lang, P., Masarie, K., 1994a. The growth rate and
distribution of atmospheric methane. J. Geophys. Res. 99, 17021 – 17043.
• Dlugokencky, E., Steele, L.P., Lang, P., Masarie, K., 1995. Atmospheric methane at
Mauna Loa and Barrow observatories: presentation and analysis of in situ measurements.
J. Geophys. Res. 100, 23103 – 23113.
• Dlugokencky, E., Masarie, K., Lang, P., Tans, P., 1998. Continuing decline in the growth
rate of the atmospheric methane burden. Nature 393, 447 – 450.
• Denier van der Gon, H., Neue, H.U., 1995. Influence of organic matter incorporation on
the methane emission from a wetland rice field. Global Biogeochem. Cycles 9, 11 – 22.
• Dobbie, K., Smith, K., 1996. Comparison of CH4 oxidation rates in woodland, arable and
set aside soils. Soil Biol. Biochem. 28, 1357 – 1365.
• Etheridge, D. M., Steele, L., Francey, R. J., & Langenfelds, R. L. (1998). Atmospheric
methane between 1000 AD and present: Evidence of anthropogenic emissions and
climatic variability. Journal of Geophysical Research: Atmospheres, 103(D13), 15979-
15993.
• Etheridge, D. M., Pearman, G. I., & Fraser, P. J. (1992). Changes in tropospheric
methane between 1841 and 1978 from a high accumulation‐rate Antarctic ice core. Tellus
B, 44(4), 282-294.
• Etminan, M., Myhre, G., Highwood, E. J., & Shine, K. P. (2016). Radiative forcing of
carbon dioxide, methane, and nitrous oxide: A significant revision of the methane
radiative forcing. Geophysical Research Letters, 43(24), 12-614.
• Fiore, A. M. (2002). Background ozone over the United States in summer: Origin, trend,
and contribution to pollution episodes. Journal of Geophysical Research, 107(D15).
• Fuglestvedt, J., Jonson, J., Isaksen, I., 1994. Effects of reductions in stratospheric ozone
on tropospheric chemistry through changes in photolysis rates. Tellus 46B, 172 – 192.
• Florez-Leiva, L., Damm, E., & Farías, L. (2013). Methane production induced by
dimethylsulfide in surface water of an upwelling ecosystem. Progress in
oceanography, 112, 38-48.
• Fishman, J., 1985. Ozone in the troposphere. In: Whitten, R.C., Prasad, S.S. (Eds.),
Ozone in the Free Atmosphere. Van Nostrand-Reinhold, New York, NY, pp. 161 – 194.
• Fishman, J., Solomon, S., Crutzen, P.J., 1979. Observational and theoretical evidence in
support of a significant in-situ photochemical source of tropospheric ozone. Tellus 31,
432 – 446.
• Gettelman, A., Holton, J.R., Rosenlof, K.H., 1997. Mass fluxes of O3, CH4, N2O and
CF2Cl2 in the lower atmosphere calculated from observational data. J. Geophys. Res.
102, 19149 – 19159.
• Huang, Y., Sass, R.L., Fisher, F.M., 1997. Methane emission from Texas rice paddy
soils: 1. Quantitative multi-year dependence of CH4 emission on soil, cultivar and grain
yield. Global Change Biol. 3, 479 – 489.
• Hudgens, D., Yavitt, J., 1997. Land-use effects on soil methane and carbon dioxide fluxes
in forests near Ithaca, New York. Ecoscience 4, 214 – 222
• Harper, L.A., Denmead, O.T., Freney, J.R., Byers, F.M., 1999. Direct measurements of
methane emissions from grazing and feedlot cattle. J. Anim. Sci. 77, 1392 – 1401.
• Intergovernmental Panel on Climate Change, 1995. In: Houghton, J.T., Meira Filho, L.,
Bruce, J., Lee, H., Callander, B., Haites, E., Harris, H., Maskell, K. (Eds.), Climate
Change 1994. Cambridge Univ. Press, Cambridge, UK, 339 pp.
• Intergovernmental Panel on Climate Change, 1996. In: Houghton, J.T., Meira Filho,
L.G., Callander, B.A., Harris, N., Kattenberg, A., Maskell, K. (Eds.), Climate Change
1995: The Science of Climate Change. Cambridge Univ. Press, Cambridge, UK, 572 pp.
• Isaksen, I.S.A., 1988. Is the oxidizing capacity of the atmosphere changing? In: Rowland,
F.S., Isaksen, I.S.A. (Eds.), The Changing Atmosphere. Wiley, New York, NY, pp. 141 –
157.
• Jouzel, J., Jouzel, N.I., Barkov, J.M., Barnola, M., Bender, J., Chappellaz, C., Genthon,
V.M., Kotlyakov, V., Lipenkov, C., Lorius, J.R., Petit, D., Raynaud, G., Raisbeck, C.,
Ritz, T., Sowers, M., Stievenard, F., Yiou, F., Yiou, P., 1993. Extending the Vostok ice-
core record of palaeoclimate to the penultimate glacial period. Nature 364, 407 – 412.
• Judd, A. G., Hovland, M., Dimitrov, L. I., Garcia Gil, S., & Jukes, V. (2002). The
geological methane budget at continental margins and its influence on climate
change. Geofluids, 2(2), 109-126.
• Joabsson, A., Christensen, T.R., Wallen, B., 1999. Vascular plant controls on methane
emissions from northern peatforming wetlands. Trends Ecol. Evol. 14, 385 – 388.
• Khalil, M., Rasmussen, R., 1987. Atmospheric methane: trends over the last 10,000
years. Atmos. Environ. 21, 2445 – 2452.
• Khalil, M., Rasmussen, R., Moraes, F., 1993a. Atmospheric methane at Cape Meares:
analysis of a high-resolution data base and its environmental implications. J. Geophys.
Res. 98, 14753 – 14770.
• Khalil, M., Rasmussen, R., Shearer, M., Ge, S., Rau, J., 1993b. Methane from coal
burning. Chemosphere 26, 473 – 477.
• Khalil, M., Rasmussen, R., 1994a. Global emissions of methane during the last several
centuries. Chemosphere 29, 833 – 842.
• Khalil, M., Rasmussen, R., 1994b. Trends in atmospheric methane. Pure and Applied
Chemistry. 66, Special Report: Methane in the Atmosphere, Commission on Atmospheric
Chemistry.
• Khalil, M.A.K., Rasmussen, R.A., 1985. Causes of increasing atmospheric methane:
depletion of hydroxyl radicals and the rise of emissions. Atmos. Environ. 13, 397 – 407.
• Khalil, M., Rasmussen, R., 1987. Atmospheric methane: trends over the last 10,000
years. Atmos. Environ. 21, 2445 – 2452.
• Khalil, M., Rasmussen, R., Shearer, M., Dalluge, R., Ren, L., Duan, C., 1998. Factors
affecting methane emissions from rice fields. J. Geophys. Res. 103, 25219 – 25231.
• King, J.Y., Reeburgh, W.S., Regli, S.K., 1998. Methane emission and transport by arctic
sedges in Alaska: results of a vegetation removal experiment. J. Geophys. Res. 103,
29029 – 29083.
• Kirchgessner, D.A., 2000. Fossil fuel industries. In: Khalil, M. (Ed.), Atmospheric
Methane: Its Role in the Global Environment. Springer-Verlag, New York, NY, pp. 263 –
279.
• Koneswaran G, Nierenberg D. 2008. Global farm animal production and global warming:
impacting and mitigating climate change. Environ Health Perspect
.116:578–82.
• Legrand, M. R., Lorius, C., Barkov, N. I., & Petrov, V. N. (1988). Vostok (Antarctica)
ice core: Atmospheric chemistry changes over the last climatic cycle (160,000
years). Atmospheric Environment (1967), 22(2), 317-331.
• Lelieveld, J., Crutzen, P., Dentener, F., 1998. Changing concentration, lifetime and
climate forcing of atmospheric methane. Tellus 50B, 128 – 150.
• Lelieveld, J., Crutzen, P., Bruhl, C., 1993. Climate effects of atmospheric methane.
Chemosphere 26, 739 – 767.
• Minami, K., Neue, H.U., 1994. Rice paddies as a methane source. Clim. Change 27, 13 –
26.
• Minami, K., 1997. Atmospheric methane and nitrous oxide: sources, sinks and strategies
for reducing agricultural emissions. Nutr. Cycling Agroecosyst. 49, 203 – 211.
• Mitra, S., Jain, M., Kumar, S., Bandyopadhyay, S., Kalra, N., 1999. Effect of rice
cultivars on methane emission. Agricult., Ecosyst. Environ. 73, 177 – 183.
• Mancinelli, R., 1995. The regulation of methane oxidation in soil. Annu. Rev. Microbiol.
49, 581 – 605.
• Mosier, A., Delgado, J., Cochran, V., Valentine, D., Parton, W., 1997a. Impact of
agriculture on soil consumption of atmospheric CH4 and a comparison of CH4 and N2O
flux in subarctic, temperature and tropical grasslands. Nutr. Cycling Agroecosyst. 49, 71
– 83.
• Mosier, A., Parton, W., Valentine, D., Ojima, D., Schimel, D., Heinemeyer, O., 1997b.
CH4 and N2O fluxes in the Colorado shortgrass steppe: 2. Long-term impact of land use
change. Global Biogeochem. Cycle 11, 29 – 42.
• Nakazawa, T., Machida, T., Tanaka, M., Fujii, Y., Aoki, S., & Watanabe, O. (1993).
Differences of the atmospheric CH4 concentration between the Arctic and Antarctic
regions in pre‐industrial/pre‐agricultural era. Geophysical Research Letters, 20(10), 943-
946.
• Nisbet, E. G., Dlugokencky, E. J., Manning, M. R., Lowry, D., Fisher, R. E., France, J.
L., ... & Ganesan, A. L. (2016). Rising atmospheric methane: 2007–2014 growth and
isotopic shift. Global Biogeochemical Cycles, 30(9), 1356-1370.
• Neue, H.-U., Roger, P.A., 2000. Rice agriculture: factors controlling emissions. In:
Khalil, M. (Ed.), Atmospheric Methane: Its Role in the Global Environment. Springer-
Verlag, New York, NY, pp. 134 – 169.
• Osborn, T., Wigley, T., 1994. A simple model for estimating methane concentration and
lifetime variations. Clim. Dyn. 9, 181 – 193.
• Peer, R., Thorneloe, S., Epperson, D., 1993. A comparison of methods for estimating
global methane emissions from landfills. Chemosphere 26, 387 – 400.
• Petrenko, V. V., Smith, A. M., Schaefer, H., Riedel, K., Brook, E., Baggenstos, D., ... &
Severinghaus, J. P. (2017). Minimal geological methane emissions during the Younger
Dryas–Preboreal abrupt warming event. Nature, 548(7668), 443-446.
• Petit, J. R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J. M., Basile, I., ... &
Stievenard, M. (1999). Climate and atmospheric history of the past 420,000 years from
the Vostok ice core, Antarctica. Nature, 399(6735), 429-436.
• Prinn, R.G., Weiss, R.F., Miller, B.R., Huang, J., Alyea, F.N., Cunnold, D.M., Fraser,
P.J., Hartley, D.E., Simmonds, P.G., 1995. Atmospheric trends and lifetime of CH3CCl3
and global OH concentrations. Science 269, 187 – 198.
• Powlson, D., Goulding, K., Willison, T., Webster, C., Hutsch, B., 1997. The effect of
agriculture on methane oxidation in soil. Nutr. Cycling Agroecosyst. 49, 59 – 70.
• Prieme, A., Christensen, S., Dobbie, K., Smith, K., 1997. Slow increase in rate of
methane oxidation in soils with time following land use change from arable agriculture to
woodland. Soil Biol. Biochem. 29, 1269 – 1273.
• Penkett, S.A., 1988. Indications and causes of ozone increase in the troposphere. In:
Rowland, F.S., Isaksen, I.S.A. (Eds.), The Changing Atmosphere. Wiley, New York, NY,
pp. 91 – 103.
• Raynaud, D., Jouzel, J., Barnola, J., Chapellaz, J., Delmas, R., Lorius, C., 1993. The ice
record of greenhouse gases. Science 259, 926 – 934.
• Raynaud, D., Chappellaz, J., Barnola, J., Korotkevich, Y., Lorius, C., 1988. Climatic and
CH4 cycle implications of glacial –interglacial CH4 change in the Vostok ice core.
Nature 333, 655 – 657.
• Rinsland, C., Levine, J., Miles, T., 1985. Concentration of methane in the troposphere
deduced from 1951 infrared solar spectra. Nature 318, 245 – 249.
• Rasmussen, R.A., Khalil, M.A.K., 1981. Atmospheric methane (CH4): trends and
seasonal cycles. J. Geophys. Res. 86, 9826 – 9832.
• Ruppel, C. D., & Kessler, J. D. (2017). The interaction of climate change and methane
hydrates. Reviews of Geophysics, 55(1), 126-168.
• Shindell, D., Kuylenstierna, J. C. I., Vignati, E., van Dingenen, R., Amann, M., Klimont,
Z., . . . Fowler, D. (2012). Simultaneously Mitigating Near-Term Climate Change and
Improving Human Health and Food Security. Science, 335(6065), 183–189.
• Stavert, A. R., Saunois, M., Canadell, J. G., Poulter, B., Jackson, R. B., Regnier, P., ... &
Zhuang, Q. (2022). Regional trends and drivers of the global methane budget. Global
Change Biology, 28(1), 182-200.
• Schwietzke, S., Sherwood, O. A., Bruhwiler, L. M., Miller, J. B., Etiope, G.,
Dlugokencky, E. J., ... & Tans, P. P. (2016). Upward revision of global fossil fuel
methane emissions based on isotope database. Nature, 538(7623), 88-91.
• Saunois, M., Stavert, A. R., Poulter, B., Bousquet, P., Canadell, J. G., Jackson, R. B., ...
& Zhuang, Q. (2020). The global methane budget 2000–2017. Earth system science
data, 12(3), 1561-1623.
• Sass, R.L., Fisher, F.M., 1997. Methane emissions from rice paddies: a process study
summary. Nutr. Cycling Agroecosyst. 49, 119 – 127.
• Saunois, M., Bousquet, P., Poulter, B., Peregon, A., Ciais, P., Canadell, J. G.,
Dlugokencky, E. J., Etiope, G., Bastviken, D., Houweling, S., Janssens-Maenhout, G.,
Tubiello, F. N., Castaldi, S., Jackson, R. B., Alexe, M., Arora, V. K., Beerling, D. J.,
Bergamaschi, P., Blake, D. R., Brailsford, G., Brovkin, V., Bruhwiler, L., Crevoisier, C.,
Crill, P., Covey, K., Curry, C., Frankenberg, C., Gedney, N., Höglund-Isaksson, L.,
Ishizawa, M., Ito, A., Joos, F., Kim, H.-S., Kleinen, T., Krummel, P., Lamarque, J.-F.,
Langenfelds, R., Locatelli, R., Machida, T., Maksyutov, S., McDonald, K. C., Marshall,
J., Melton, J. R., Morino, I., Naik, V., O'Doherty, S., Parmentier, F.-J. W., Patra, P. K.,
Peng, C., Peng, S., Peters, G. P., Pison, I., Prigent, C., Prinn, R., Ramonet, M., Riley, W.
J., Saito, M., Santini, M., Schroeder, R., Simpson, I. J., Spahni, R., Steele, P., Takizawa,
A., Thornton, B. F., Tian, H., Tohjima, Y., Viovy, N., Voulgarakis, A., van Weele, M.,
van der Werf, G. R., Weiss, R., Wiedinmyer, C., Wilton, D. J., Wiltshire, A., Worthy, D.,
Wunch, D., Xu, X., Yoshida, Y., Zhang, B., Zhang, Z., and Zhu, Q.(2016). The global
methane budget 2000–2012. Earth System Science Data, 8(2), 697-751.
• Smith, L.K., Lewis, W.M., Chanton, J.P., Cronin, G., Hamilton, S.K., 2000. Methane
emissions from the Orinoco River floodplain, Venezuela. Biogeochemistry 51, 113 –
140.
• Steele, L., Dlugokencky, E., Lang, P., Tans, P., Martin, R., Masarie, K., 1992. Slowing
down of the global accumulation of atmospheric methane during the 1980s. Nature 358,
313 – 316.
• Subak, S., 1994. Methane from the House of Tudor and the Ming Dynasty: anthropogenic
emissions in the sixteenth century. Chemosphere 29, 843 – 854.
• Stern, D., Kaufmann, R., 1996. Estimates of global anthropogenic methane emissions
1860 – 1993. Chemosphere 33, 159 – 176.
• Štreimikienė, D., & Balezentis, T. (2016). Kaya identity for analysis of the main drivers
of GHG emissions and feasibility to implement EU “20–20–20” targets in the Baltic
States. Renewable and Sustainable Energy Reviews, 58, 1108-1113.
• Thompson, A., 1992. The oxidizing capacity of the Earth’s atmosphere: probable past
and future changes. Science 256, 1157 – 1165.
• Tahir, M. S., Shahzad, K., Shahid, Z., Sagir, M., Rehan, M., & Nizami, A. S. (2015).
Producing methane enriched biogas using solvent absorption method. Chem. Eng.
Trans, 45.
• Van den Pol-Van Dasselaar, A., Van Beusichem, M.L., Oenema, O., 1999. Methane
emissions from wet grasslands on peat soil in a nature preserve. Biogeochemistry 44, 205
– 220.
• Van der Nat, F.J., Middelburg, J.J., 2000. Methane emission from tidal freshwater
marshes. Biogeochemistry 49, 103 – 121.
• Ward, G., Doxtader, K., Miller, W., Johnson, D., 1993. Effects of intensification of
agricultural practices on emission of greenhouse gases. Chemosphere 26, 87 – 93.
• World Meteorological Organization, 1985. Scientific Assessment of Ozone Depletion:
1985. Global Ozone and Research and Monitoring Project Report 16, Geneva.
• World Meteorological Organization, 1995. Scientific Assessment of Ozone Depletion:
1994. Global Ozone and Research And Monitoring Project Report 37, Geneva.
• Wuebbles, D.J., D.R. Easterling, K. Hayhoe, T. Knutson, R.E. Kopp, J.P. Kossin, K.E.
Kunkel, A.N. LeGrande, C. Mears, W.V. Sweet, P.C. Taylor, R.S. Vose, and M.F.
Wehner, 2017: Our globally changing climate. In Climate Science Special Report: Fourth
National Climate Assessment.1:35-72.
• Wuebbles, D. J., & Hayhoe, K. (2002). Atmospheric methane and global change. Earth-
Science Reviews, 57(3-4), 177-210.
• Wang, Y., Jacob, D.J., 1998. Anthropogenic forcing on tropospheric ozone and OH since
preindustrial times. J. Geophys. Res. 103, 31123 – 31135.
• Yagi, K., Tsuruta, H., Minami, K., 1997. Possible options for mitigating methane
emission from rice cultivation. Nutr. Cycling Agroecosyst. 49, 213 – 220.
• Yavitt, J.B., Williams, C.J., Wieder, R.K., 2000. Controls on microbial production of
methane and carbon dioxide in three Sphagnum-dominated peatland ecosystems as
revealed by a reciprocal field peat transplant experiment. Geomicrobiol. J. 17, 61 – 88.

You might also like