Reviews of Geophysics - 2020 - Nisbet - Methane Mitigation Methods To Reduce Emissions On The Path To The Paris Agreement

Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

REVIEW ARTICLE Methane Mitigation: Methods to Reduce Emissions,

10.1029/2019RG000675
on the Path to the Paris Agreement
Key Points:
• The atmospheric methane burden is
E. G. Nisbet1, R. E. Fisher1, D. Lowry1, J. L. France1, G. Allen2, S. Bakkaloglu1, T. J. Broderick3,
rising fast; this growth is an M. Cain4, M. Coleman5, J. Fernandez1, G. Forster6, P. T. Griffiths7, C. P. Iverach8, B. F. J. Kelly8,
increasing threat to the Paris M. R. Manning9, P. B. R. Nisbet‐Jones1, J. A. Pyle7, A. Townsend‐Small10, A. al‐Shalaan1,
Agreement of the UN Framework
Convention on Climate Change
N. Warwick7, and G. Zazzeri11
(UNFCCC) 1
• New gas measurement technologies,
Department of Earth Sciences, Royal Holloway, University of London, Egham, UK, 2Centre for Atmospheric Science,
on the ground and in the air, have University of Manchester, Manchester, UK, 319 Jenkinson Rd., Chisipite, Harare, Zimbabwe, 4Atmospheric Oceanic and
greatly improved our ability to locate Planetary Physics, Department of Physics, University of Oxford, Oxford, UK, 5Department of Meteorology, University of
and quantify emissions and to Reading, Reading, UK, 6National Centre for Atmospheric Sciences (NCAS), School of Environmental Sciences, University
identify mitigation targets
of East Anglia, Norwich, UK, 7Department of Chemistry, Cambridge, UK, 8School of Biological, Earth and Environmental
• Emissions can be cut by ending fossil
fuel emissions, cutting biomass Science, University of New South Wales, Sydney, New South Wales, Australia, 9Climate Change Research Institute,
burning, improving landfills, School of Geography Environment and Earth Sciences, Victoria University of Wellington, Wellington, New Zealand,
10
especially in the tropics, and Department of Geology, University of Cincinnati, Cincinnati, OH, USA, 11Department of Physics and Grantham
changing cattle farming practice Institute, Imperial College London, London, UK

Correspondence to:
Abstract The atmospheric methane burden is increasing rapidly, contrary to pathways compatible
E. G. Nisbet, with the goals of the 2015 United Nations Framework Convention on Climate Change Paris Agreement.
[email protected] Urgent action is required to bring methane back to a pathway more in line with the Paris goals. Emission
reduction from “tractable” (easier to mitigate) anthropogenic sources such as the fossil fuel industries and
Citation: landfills is being much facilitated by technical advances in the past decade, which have radically improved
Nisbet, E. G., Fisher, R. E., Lowry, D., our ability to locate, identify, quantify, and reduce emissions. Measures to reduce emissions from
France, J. L., Allen, G., Bakkaloglu, S.,
“intractable” (harder to mitigate) anthropogenic sources such as agriculture and biomass burning have
et al. (2020). Methane mitigation:
methods to reduce emissions, on the received less attention and are also becoming more feasible, including removal from elevated‐methane
path to the Paris agreement. Reviews of ambient air near to sources. The wider effort to use microbiological and dietary intervention to reduce
Geophysics, 58, e2019RG000675.
emissions from cattle (and humans) is not addressed in detail in this essentially geophysical review. Though
https://doi.org/10.1029/2019RG000675
they cannot replace the need to reach “net‐zero” emissions of CO2, significant reductions in the methane
Received 13 AUG 2019 burden will ease the timescales needed to reach required CO2 reduction targets for any particular future
Accepted 10 JAN 2020 temperature limit. There is no single magic bullet, but implementation of a wide array of mitigation and
Accepted article online 14 JAN 2020
emission reduction strategies could substantially cut the global methane burden, at a cost that is relatively
low compared to the parallel and necessary measures to reduce CO2, and thereby reduce the atmospheric
methane burden back toward pathways consistent with the goals of the Paris Agreement.
Plain Language Summary Methane is a powerful climate warmer, and the amount of methane in
the air is growing rapidly. Reducing human‐caused methane emissions is urgent if the 2015 United Nations
Paris Agreement to limit climate warming is to succeed. There is hope, though the problem of methane
mitigation is very wide and complex. Much of the task is in finding, identifying, and quantifying emissions.
Rapid technical advances are making it much easier to locate and thus cut emissions from fossil fuel industries
(gas, coal, and oil). Assessing emissions from landfill and sewage facilities is also becoming easier. In
particular, poorly regulated landfills in fast‐growing tropical megacities need attention. Agricultural emissions
are less tractable but may also be reduced to some extent, especially by improving manure management.
Many methane mitigation options offer cost‐effective approaches to cut global warming and bring the
amount of methane in the air back to a pathway that is consistent with the aims of the Paris Agreement.

© 2020. The Authors. 1. Introduction


This is an open access article under the
1.1. Methane's Post‐2007 Rise and Growing Distance From the Paris Target
terms of the Creative Commons
Attribution License, which permits use, Methane is a remarkably attractive target for reducing atmospheric greenhouse warming. It is the second-
distribution and reproduction in any
medium, provided the original work is
most important anthropogenic greenhouse gas, and a major contributor to the increase in tropospheric
properly cited. ozone, also a major greenhouse gas. Methane has a 100‐yr global warming potential of 32 compared to

NISBET ET AL. 1 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

CO2 (Etminan et al., 2016); thus, removing 1 Mg (a ton) of methane has


real impact (Collins et al., 2018). Moreover, with an average tropospheric
lifetime of about 9–10 yr (Dlugokencky et al., 2011), successful moves to
reduce methane emissions can have rapid effectiveness.
Since 2007 the atmospheric methane burden has risen sharply. It began
growing by about 6 ppb/yr in 2007, after a sustained pause at the start of
the millennium. The growth rate accelerated in 2014 and is currently
around 10 ppb/yr (Nisbet et al., 2014; Nisbet et al., 2016; Nisbet et al.,
2019). Note that the words “concentration” and “level,” terms derived
from wet chemistry, are often used in popular media when discussing
methane. The gas‐specific terms “mole fraction” and “mixing ratio” are
preferred for use in this report.
The rise in atmospheric methane has been accompanied by a significant
shift in the C‐isotope ratio of atmospheric methane to values more
depleted in 13C (Nisbet et al., 2016, 2019). The post‐2007 growth and iso-
topic shift has now been sustained for over a decade, but the reasons for
the change are still not well understood. Possible explanations for this
growth and the concurrent isotopic shift (for isotopic values of sources,
see Sherwood et al., 2017); may include increases in biogenic emissions
(especially in the tropics and subtropics), changes in the chemical sinks
of methane by atmospheric OH or Cl, increased fossil fuel emissions
coupled with declining biomass burning (Nisbet et al., 2019, 2016;
Schaefer et al., 2016; Schwietzke, Sherwood et al., 2017; Turner et al.,
2017, 2019; Rigby et al., 2017; Worden et al., 2017), more oxidation of
methane by methanotrophy in forest soils (Ni & Groffman, 2018), or,
Figure 1. Methane and the Paris Agreement. Top panel shows evolution of
the mean global atmospheric mixing ratio (green open circles) compared more likely, some combination of all these factors.
to Representative Concentration Pathway RCP2.6 (Collins et al., 2013) as One illustrative pathway that would lead to compliance with the 2015
used in the last IPCC Assessment and consistent with the Paris Agreement.
United Nations (UN) Paris Agreement of the United Nations Framework
RCP2.6 concentrations peaked in 2012, whereas those for the RCP 4.5
(dashed) pathway peak around 2045, while RCP 8.5 (dotted) mole fractions Convention on Climate Change (UNFCCC, 2015) is Representative
Concentration Pathway 2.6 (RCP 2.6) (Collins et al., 2013; Rogelj et al.,
continue to rise throughout the 21st century. Lower panel shows differences
in radiative forcing from the RCP2.6 pathway and actual evolution of 2012) (Figure 1). This pathway envisaged an immediate and significant fall
methane (green open circles), CO2 (red circles), and nitrous oxide (blue in methane, allowing time for progress on the more difficult task of redu-
circles). Dashed lines show RCP 4.5 relative to RCP 2.6. Radiative forcing
cing CO2. This is not a unique pathway to achievement of the Paris goals,
used here is from Etminan et al. (2016), and the figure is updated from that
in Nisbet et al. (2019) using NOAA mole fraction data to June 2019. but being Paris compliant is thus commonly cited as an exemplar of the
reduction pathways needed. Yet, by 2019, as a result of the unexpected rise,
the atmospheric methane mixing ratio was over 100 ppb above the RCP 2.6 path (Nisbet et al., 2019). If added
to the atmosphere as a single pulse, this would equate to an input of about 300 Tg of methane (1 ppb is glob-
ally equivalent to about 2.8 Tg CH4).
The amount of methane in the atmosphere above preindustrial levels directly affects the allowable cumula-
tive carbon emissions budget (i.e., the total anthropogenic CO2 emitted before reaching a temperature
threshold, usually 1.5 or 2 °C). The rise thus has major implications for the UN Paris Agreement's goal to
limit climate warming to 2 °C and especially to the 2018 Intergovernmental Panel on Climate Change
(IPCC) advice to limit warming to 1.5 °C (IPCC, 2018). The global warming consequences of this rise, rela-
tive to the Paris‐compliant path, are shown in Figure 1 from Nisbet et al. (2019). Moreover, the unexpected
growth in methane has already significantly negated the expected impact of progress in controlling CO2
emissions (Nisbet et al., 2019). Haustein et al. (2017) suggest that methane is contributing to the accelerating
global warming rate at a time when CO2 emissions may be stabilizing. If growth in the methane burden con-
tinues at current rates in the coming decades, it will become impossible to meet the Paris target, especially in
view of the recent upward reevaluation of the global warming potential of methane (Etminan et al., 2016).
The Paris Agreement, drafted in 2015 and effective in 2016, is clear in its call for better technology to reduce
methane emissions. Article 10.2 and 10.5 of the agreement (UNFCCC, 2015) states that

NISBET ET AL. 2 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

“Parties, noting the importance of technology for the implementation of mitigation and adaptation actions
under this Agreement and recognizing existing technology deployment and dissemination efforts, shall
strengthen cooperative action on technology development and transfer.” And that “Accelerating, encouraging
and enabling innovation is critical for an effective, long‐term global response to climate change and promoting
economic growth and sustainable development.”
In Article 13.7, the Agreement places requirements on quantification of emissions:
“Each Party shall regularly provide the following information:
(a) A national inventory report of anthropogenic emissions by sources and removals by sinks of greenhouse
gases, prepared using good practice methodologies accepted by the Intergovernmental Panel on Climate
Change and agreed upon by the Conference of the Parties serving as the meeting of the Parties to this
Agreement; and
(b) Information necessary to track progress made in implementing and achieving its nationally determined con-
tribution under Article 4.”
There is thus a requirement for signatory nations to act effectively and urgently to reduce methane emissions.

1.2. The Challenge—Low‐Hanging Fruit or Tough Target?


Cutting methane emissions has long been identified as a low‐hanging fruit for climate action (Hansen et al.,
2000; UNEP, 2018). Shindell et al. (2017) considered as much as 110 Tg CH4 per year of abatement was pos-
sible by scaling up existing technology and industrial best practice and policy options, to provide societal
benefits that outweigh implementation cost. A significant cut in anthropogenic methane emissions would
be sufficient to reduce the overall methane burden within decades or less, to thus return to the RCP2.6 path-
way. There is wide benefit in reducing methane emissions (Boucher & Folberth, 2010) including major
reduction opportunities at less than zero cost (Warner et al., 2015).
Globally, the rate of growth in fossil fuel methane emissions may be slowing (Schaefer et al., 2016). Nisbet
et al. (2016, 2019) used isotopic evidence to infer that fossil fuel emissions have declined as a proportion
of the total methane budget, though they could not rule out an increase in absolute terms. But it is also pos-
sible (within the uncertainties of the observations) that a decline in biomass burning may be masking a rise
in fossil fuel emissions (Worden et al., 2017).
Many jurisdictions have brought in legislation to control methane emissions (Iacobuta et al., 2018). In 2019,
the Net Zero report of the U.K.'s Committee on Climate Change led to a national commitment to reduce net
national greenhouse emissions to zero by 2050 (UK CCC, 2019). New Zealand has passed a similar promise.
Though these commitments focus on CO2, most nations also intend to reduce methane emissions in addition
to cutting CO2 emissions. The focus has been on point sources where very high mole fractions of methane
can be measured in the nearby air—such as gas leaks, or gas emissions around landfills. Such sources can
typically be detected easily, precisely located and then mitigated.
For example, Canada has committed to reduce methane emissions from the oil and gas sector by 40–45% by
2025 relative to 2012 levels (Canada, 2016). To take another example, in the United Kingdom where landfill
emissions have been vigorously reduced, the national atmospheric emissions inventory suggests methane
emissions have dropped by 61% since 1990 (UK NAEI, 2019; UK NIR, 2019), a pattern seen in many devel-
oped nations. In these specific sectors, there has been much progress in abatement of emissions, at relatively
low cost by improved industrial practices to detect and prevent leaks, or to capture and use methane before it
is emitted to atmosphere. But for many sources with more diffuse emissions that are less easily detected (as
local air has low incremental methane over background), there has been little progress.
A challenge to policy‐making is that there is a major discrepancy (Leip et al., 2018; Saunois et al., 2016, 2019)
between “top‐down” estimates (Nisbet & Weiss, 2010) of the annual global methane emission, assessed from
measuring the atmosphere, and “bottom‐up” totals summing emissions estimates based on national data
(e.g., number of cows and area of rice fields): the “bottom‐up” numbers are typically much higher (e.g.
EDGAR‐ European Commission, Joint Research Centre / Netherlands Environmental Assessment
Agency, 2011). Top‐down results can cause major revisions to previous emission inventories, for example
after the work of Bergamaschi et al. (2010). Press reports of an earlier study by this team show that it

NISBET ET AL. 3 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Table 1 caused substantial revisions (https://www.theguardian.com/environ-


Estimates of Global Annual CH4 Budget in 2012 (Range) (Tg/yr; Saunois ment/2006/jun/22/climatechange.climatechangeenvironment). Chang
et al., 2016; See Also Saunois et al., 2019 for Slightly Different Updates)
et al. (2019) also illustrate this problem: they found, compared to previous
Sources Top‐down Bottom‐up estimates, a larger than expected contribution of ruminant enteric emis-
Natural sources total 221 (192–302) 386 (259–532) sions to the increasing trend in global methane emissions between 2000
Natural wetlands 172 (155–201) 187 (155–235) and 2012. This discrepancy may point to errors in either or both top‐down
Other Natural sources 49 (22–137) and bottom‐up assessment methodologies, a problem that does not instill
Agriculture and waste 200 (122–213) 197 (183–211)
confidence in using bottom‐up emissions estimates to identify the opti-
Biomass burning, etc. 35 (28–51) 30 (25–36)
Fossil fuels 112 (90–137) 134 (123–141) mum (least cost, most benefit) targets for reduction efforts.
Sum of sinks 555 The global challenge is clear from Table 1. The need is to cut emissions
from fossil fuels and from biomass and biofuel burning and reduce the
methane footprint of agriculture and waste. But within this global envelope, each nation has its own spec-
trum of emissions. Figure 2 illustrates the impact of a sustained effort to reduce U.K. national emissions.
Earlier emission cuts were very significant, but recently, the reduction trend has slowed or ceased. In
2017, the National Atmospheric Emissions Inventory (UK NAEI) estimated total U.K. emissions from waste
as about 756 kilotons (kt) (quoted in tons by NAEI with no errors given: 1 kt = 0.001 Tg), fugitive emissions
from energy fuels as 217 kt, 63.5 kt from combustion, and 187 kt from agriculture. If emissions from waste,
fuels, and combustion were removed, the national total could be cut a further 80% or more.
Reducing emissions from ongoing anthropogenic activities such as fossil fuel production and intensive rumi-
nant farming is essentially palliative: it helps but does not cure. More substantial reductions can be achieved
by ending or substantially reducing the activities themselves, for example, by shifting entirely away from
coal and gas as energy sources and by reducing ruminant emissions. But such actions require political debate
and major social adjustments, which are outside the scope of this journal. The more limited mitigation
options discussed here are less contentious and can be implemented within the framework of existing
commitments to the UNFCCC Paris Agreement.

2. Methodology
2.1. General Methodology
We need to develop cost‐effective methodologies for cutting emissions and, where they cannot be cut, for
removing the emitted methane from air with elevated methane around the sources. Methane reduction is
a technical challenge demanding source detection and identification, source‐specific emission flux quantifi-
cation, and effective reduction methodologies and targeting.
The task of reducing the methane burden is discussed in the following sections:
• Section 2: Location of emissions and Identification of sources,
• Section 3: Quantification of emissions,
• Section 4: Practical emission reduction—tractable emissions,
• Section 5: Emission removal—intractable emissions,
• Section 6: Reducing agricultural methane production,
• Section 7: Summary of priorities,
• Section 8: Outlook.
Some of these tasks are straightforward but most are not simple. For example, location of emissions, espe-
cially major point sources, should intuitively be expected to be an easy task but in practice may be complex
as what initially appears to be a point source may actually comprise many small subsources. For example, in
a gasfield, leakage from producing wells may be relatively minor, while major emissions may be discovered
from unexpected sites such as water processing facilities (e.g., Iverach et al., 2015; Kelly et al., 2015;
Schwietzke, Pétron et al., 2017), or forgotten old (legacy) uncapped exploration wells (Day et al., 2015).
Source identification can be a complex task: For example, very different major sources can be closely colo-
cated, such as cattle feedlots within gasfields, or urban landfills near biodigesters, sewage facilities, and
gas distribution plants. Industrial area planning and land use often means that such facilities are located
near to one another. Urban source types may also be multiple, yet geographically close. Attribution of emis-
sions to specific sources thus can be a challenging task, demanding high‐precision measurement of methane

NISBET ET AL. 4 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

and also winds and boundary layer mixing heights, and often isotopic ana-
lysis as well as measurement of other proxy tracer species such as ethane
or carbon dioxide (Allen, 2016).
Accurate quantification of emissions is necessary if the most cost‐effective
reduction strategies are to be targeted, but quantification of emission flux
in plumes remains an imprecise art. Kang et al. (2016) showed the dispro-
portionate impact of major gasfield leaks—the so‐called “superemitters.”
In the United States Omara et al. (2018) found the top 5% of sites
accounted for 50% of cumulative emissions. Identifying, quantifying,
and stopping these superleaks is the first step in emission reduction.
More generally, jurisdictions need to quantify, declare, and verify emis-
sion declarations at all levels, from local emitter to nation state, but this
remains very limited outside Europe (e.g., see Bergamaschi et al., 2018)
and North America.
Even if major emissions are correctly identified, located, and quantified,
then targeting the most cost‐effective ways to reduce emissions is not sim-
ple. For example, Kuwait has a major hydrocarbon industry, but field
campaigns carried out there by the Royal Holloway (RHUL) group there
suggest, counterintuitively, that cutting Kuwait's landfill methane output
(Figure 3) may be a cost‐effective first‐choice target for cutting overall
national emissions (al‐Shalaan, 2019), in part perhaps because Kuwait's
oil industry already operates a program for emissions reduction.
Removal of emissions from the air, after they have been emitted, is a major
part of the CO2 carbon capture and storage discussion but has been little
discussed for methane. This enquiry is a task that deserves attention and
technological development.
In this synopsis for Reviews of Geophysics we focus on essentially geophy-
Figure 2. U.K. national inventory of methane emissions by source sector. sical methods to reduce emissions. Sustained reduction of the atmo-
Note the significance of the decline in emissions from waste and energy spheric methane burden will almost certainly demand much wider
sources in the late 1990s, but the relative lack of progress in recent years.
actions, both in management of ruminants and also perhaps societal
From UK NAEI (2019).
change including demand‐reduction for food derived from ruminants.
There is a broad literature on these major topics. The focus here is tactical: what can be done now. The wider
strategic issues are briefly discussed later (section 8).

2.2. Locating Emissions


Rapid technical progress is being made, both in pinpointing emissions more accurately and quickly, and in
finding smaller point sources. Various vehicle, drone, and aircraft‐mounted technologies can be used.
Although most of the simpler current technologies are still only able to provide rudimentary information
on flux quantification, they are generally effective in finding leaks and suitable for localization of emission
sources (Ravikumar et al., 2019).
Much progress is being made in leak detection and repair. The U.S. ARPA‐E‐MONITOR program (https://
arpa‐e.energy.gov/?q=program‐projects/MONITOR) has the ambitious goal to “detect and measure methane
leaks as small as 1 ton per year from a site 10m x 10m in area with a certainty that would allow 90% reduction in
methane loss for an annual site cost of $3,000.” [1 U.S. ton = 907.1 kg]. If this is achieved, it will radically
improve our ability to cut emissions not only in the United States but worldwide. Currently, much leak detec-
tion in gasfields is by optical gas imaging using passive infrared cameras. For example, over the Bakken
Formation in the Williston Basin of North Dakota, USA, Englander et al. (2018) used helicopter‐borne infra-
red optical gas imaging to survey about 1,000 well pads, in conjunction with use of a light aircraft to quantify
fluxes in a randomly selected sample of 33 plumes. However, the effectiveness of cameras depends on the
backdrop, and their usefulness declines with distance, a problem because it limits their usefulness in surprise
near‐ but off‐site drive‐by inspections by regulatory bodies. Cameras are effective close to large known
sources, such as superemitters, and are excellent for detecting large leaks at gasfields from distances of 10–

NISBET ET AL. 5 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

13
Figure 3. Kuwait landfill; with δ CCH4 of −57‰, landfills may produce up to half of Kuwait's methane emissions (al‐Shalaan, 2019). Photo: D. Lowry.

100 m (Ravikumar et al., 2016) and for big emissions over leaking street gas mains, but cameras are less able
to find abundant smaller leaks in gas collection, transport and urban distribution networks.
In contrast to camera detection, mobile instruments that map by optical spectroscopy are much more sensi-
tive. They can detect smaller leaks, with good precision and flux quantification (Weller et al., 2018) down-
wind from suspected sources. In the past decade, cavity‐based instruments have radically improved the
precision and mobility of in‐the‐field measurement of methane and other associated species (Baer et al.,
2012; Caulton et al., 2018; Crosson, 2008). In particular, compared to older gas chromatography, with its
low portability and calibration requirements, modern cavity‐enhanced absorption spectrometry instruments
have much better precision, robust portability, and more stable calibration in challenging environments
(von Fischer et al., 2017; Zazzeri et al., 2015). Other techniques are also available for methane measure-
ment—for example quantum cascade lasers and open path analyzers (McDermitt et al., 2011; Nelson
et al., 2004). Other species such as ethane can be measured simultaneously, to differentiate between gas
supply leaks and sewer emissions. Isotopic analysis can also be used to determine the nature of the source
(e.g., gas leaks or sewage pipes) (Fries et al., 2018; Lowry et al., 2019, 2009; Phillips et al., 2013).
2.2.1. Location by Vehicle‐ and Boat‐Mounted Mapping
It is becoming increasingly inexpensive and rapid to monitor leaks in both production gasfields and urban
reticulation systems by routine drive‐by surveys. Rapid precise mapping of methane emissions is now
increasingly possible with offsite systems coupled to real‐time global positioning system (GPS) location
(Bamberger et al., 2014; Phillips et al., 2013, Rella et al., 2015, von Fischer et al., 2017). Such instruments
can be deployed at remote places, and can be mounted on small domestic SUV vehicles (Zavala‐Araiza
et al., 2017; Zazzeri et al., 2015). Vehicle surveys can be supplemented by flying small unmanned aerial
vehicles (UAVs) equipped with lightweight sensors (see section 2.2.2 below).
Continuously measuring GPS‐linked mobile systems operate with anemometers to measure wind directions
and speeds. As roads typically occur within 100 m to 1 km upwind and downwind of industrial, landfill, and
cattle feedlot sources, this allows accurate location of emissions. Thus, in many cases biogas and broad land-
fill emission peaks can be separated by plume shape and wind direction analysis. In many cases, the likely
source can immediately be pinpointed. Figure 4 illustrates a typical vehicle‐mounted system. The costs of
vehicle‐mounted campaigns are low. The instrument and associated equipment costs can be below U.S.
$100,000. Common SUV vehicles can easily be adapted for the task. They can also deploy air hoses on
pop‐up balloons, to take air samples (e.g., for isotopic analysis) (Steiger et al., 2015).
The instruments map out methane mole fraction in the ambient air around the vehicle or above the vehicle's
roof. Assuming a steady background air flow, the local increment over background can then be measured
while the vehicle moves at speeds that can be up to ~80 km/hr, though typically optimized for plume capture
at 20–30 km/hr. Methane peaks identify emission sources: then the sources can be located and mapped care-
fully using local access roads. Similarly, boats can be used to map emissions from offshore gas platforms
(Riddick et al., 2019), though here the problem is that the emissions may be from locations such as offshore
platform flares that are above the height of the boat's mast intake.

NISBET ET AL. 6 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Helfter et al. (2019) used measurements from a North Sea ferry to estimate methane emissions from the
United Kingdom and Irish republic. They mounted a cavity‐ringdown analyzer on the bow of a commercial
freight ferry routinely sailing between Scotland and Belgium and compared the results with measurements
made at Mace Head on the west coast of Ireland, using a mass balance approach. They found that estimated
annual methane emission from the study region was 2.55 ± 0.48 Tg. This result is comparable with the
2.29 Tg/yr total of estimates submitted to the UNFCCC.
Mapping by mobile analyzer is rapid, and field campaign costs can be modest, apart from the capital costs of
the instrument and vehicle. In the Royal Holloway laboratory, located SW of the London urban area, Greater
London's leaks are being mapped in the duration of a PhD thesis. Hong Kong sources were mapped within a
5‐day working week during a period of steady methane background set by prevailing marine SE Monsoon
winds, arriving from the South China Sea. Such campaigns are not in themselves quantitative, nor do they
characterize sources as the surveys merely locate major “hot spots”. But in most cases that is enough.
Emissions sources can be pinpointed quickly and then remediation or mitigation can be considered. In some
cases, halting emissions can be profitable (e.g., from broken pipes or leaking production wells). However, in
other cases, where leaks do not pose a significant safety risk, repair may have low priority. In such cases,
regulatory intervention may be needed: Thus, there is a strong argument that regulatory authorities should
have the capacity for independent discovery of leaks and the power to impose financial penalties.
2.2.2. Location by Drone (UAV) Surveys.
Drones (UAVs) offer powerful opportunities for rapid identification and hence reduction of concentrated
methane leaks (Fox et al., 2019). Aerial measurement and sampling have the potential to map out emission
plumes in three spatial dimensions, as well as time. Moreover, if the drones carry anemometers, eddy flux
calculations may be carried out to determine emission fluxes.
Small drones can carry instruments to map methane plumes close to sources and use coincident wind mea-
surements to derive flux using a range of approaches including Gaussian plume modeling and mass balan-
cing. Already it is possible to use these inexpensive light sensors to patrol gas installations. For typical small
quadcopter and octocopter drones, the lifting ability is sufficient to carry small sensors able to achieve
methane measurement to moderate precision. Most low‐cost low‐weight sensors are capable of measuring
methane to 2‐ppm precision (Collier‐Oxandale et al., 2018; Shah et al., 2019), which should be capable of
detecting significant emissions around known candidate sites. However the precision and accuracy of light-
weight small sensors is currently poor compared to vehicle‐mounted or static ground‐mounted instruments.
Larger UAVs using inexpensive lasers in the near infrared can achieve better precision, to ~100 ppb.
Recently, more expensive midinfrared lasers have been used, exploiting methane's strong absorption band
around 3.3 μm, and have demonstrated performance in the 1‐ to 10‐ppb level (Golston et al., 2017; Shah
et al., 2019) when mounted on drones weighing 5 kg or less. The technology is changing rapidly. Better than
10 ppb (at 1 Hz) precision instrumentation is now available and in use on commercially available UAV plat-
forms (Shah et al., 2019), and 1 ppm precision is likely soon. Further marked improvement in the precision,
price, and availability of UAV‐mounted sensors should be expected in the near future. Small drone‐carried
sensors have much promise for safe close‐up mapping of air with high methane mole fractions around major
leaks, for example, in mapping the 3‐D distribution of a plume from a leaking gas installation, and hence
quantifying the flux (Emran et al., 2017).
UAV surveys are useful in large‐scale screening of gas infrastructure for major point sources (e.g., over a
spread‐out gasfield or landfill site) to find and quantify large leaks quickly (Allen et al., 2019; Barchyn
et al., 2017; Yang et al., 2018). Many leaks are episodic “one‐offs”—the result of errors such as faulty or for-
gotten valves, ill‐fitted connections or equipment failure. Sustained drone patrols, carrying low‐cost sensors
close to known gas systems, can identify these transient events quickly and cheaply. Such systems have great
potential and are likely soon to become a high priority for industry and regulatory bodies (e.g., by the U.K.
Environment Agency, see Allen et al., 2018). As equipment costs are low, it should be possible and inexpen-
sive to ensure each gas installation has frequent (daily or potentially multiple times a day) routine inspection
by automated drones. This may even be profitable, both in terms of less leakage and in safety improvements,
averting the risk of major problems and reducing insurance costs.
For isotopic measurements, drones can be used to take grab samples of air in Tedlar or aluminum foil bags
(Figure 5) to be analyzed in a laboratory. These bags only weigh a few grams and can be carried by small

NISBET ET AL. 7 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 4. (left) Schematic set up of the Royal Holloway Picarro mobile measurement system (from Zazzeri et al., 2015).
(right upper and lower) Vehicle mounting. (right) Deployment at Bacton gas terminal, Norfolk, UK.

quadcopters. Aircores, which are air samples collected in coiled tubes for later analysis in the laboratory, can
also be used to sample air with UAVs (Andersen et al., 2018).
On Ascension Island in the equatorial South Atlantic, Brownlow et al. (2017) and Greatwood et al. (2017)
report bag‐sampling air from as high as 2,700 m in the midtroposphere, well above the height of the local
South Atlantic Trade Wind Inversion at about 1,200–1,500 m that defines the top of the marine boundary
trade winds. This allowed Brownlow et al. (2017) to sample air that had traveled from the otherwise
difficult‐to‐access Congo basin in equatorial Africa, a potential future tool for validating national and regio-
nal emissions inventories.
In most locations, UAVs are permitted a maximum altitude of 100–200 m, confining them within the lower
part of the boundary layer. This is sufficient for proximal surveys on sources, but in most countries higher
altitudes are barred to drones and can only be accessed by aircraft. It should be noted however that in many
tropical nations UAV restrictions are extremely severe and in some cases wholly barred, particularly near
military bases. Moreover, in nations like the United Kingdom legal frameworks for UAV operation are
increasingly complex and obtaining permits may become challenging to small academic teams, especially
close to sensitive installations such as gas industry facilities, thus may limit the usefulness of UAV techni-
ques for leak studies, and especially the development of novel applications by university research.
2.2.3. Location by Aircraft Surveys.
Piloted aircraft capable of carrying high‐precision spectroscopic instruments are still needed for precise
measurement of regional fluxes and major emissions sources. To assess emissions from further afield, where
mixing with ambient air has taken place and the incremental methane signal is small, for example, some
kilometers downstream from a coal mine or large cattle feedlot, or to characterize bulk emissions from a city
or a major gasfield, high‐precision optical cavity‐based fast methane analyzers are needed. Currently,
instrument and battery packages typically weigh >10 kg for mole fraction precision better than 1 ppb, and
the required battery capacity for a useful survey is high (of the order of 100 amp‐hour). These are too heavy
to fly on cheap drones and require commercially available UAVs with a takeoff weights typically >20 kg.
Moreover, as mentioned above, drones typically have a height limit in most jurisdictions, depending on
airspace classification.
Karion et al. (2013, 2015) and Caulton et al. (2014) used aircraft‐based measurements to map out emissions
from unconventional oil and gas fields in the United States, while Peischl et al. (2015, 2018) and Schwietzke
et al. (2017) used aircraft equipped with a high‐precision analyzer to study several major shale gas fields.

NISBET ET AL. 8 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

13
Figure 5. Example of drone‐carried Tedlar bag sampling for methane mole fraction and δ CCH4 (Brownlow et al., 2017;
Greatwood et al., 2017).

Similarly, in the Netherlands, Yacovitch et al. (2018) used a light aircraft equipped with a high‐precision gas
analyzer and flask sampling capability to determine emissions from the giant Groningen gasfield and the
heavily agricultural surrounding area. They demonstrated that production volume from a specific produc-
tion site is not a good guide to emissions from that site and that applying inventory emission factors (i.e.,
production‐based weighting) does a poor job in assessing emissions.
Aircraft are useful in measuring agricultural emissions also. Using a light aircraft flying at 50–500 m above
ground, Hiller et al. (2014) tested the Swiss methane emissions inventory over an agricultural valley in rural
Switzerland. They measured methane mole fractions during the flights and then used two approaches, eddy
flux covariance and boundary layer budgeting, to assess fluxes. Flux estimates were higher than the Swiss
national inventory, suggesting Swiss agricultural emissions had been underestimated. Studying a powerful,
more focused source, Hacker et al. (2016) used a low flying aircraft over an Australian feedlot with 17,000
cattle. The aircraft was equipped with quantum cascade laser analyzers, and high resolution wind
turbulence measurement. Precision was adequate to apportion emissions qualitatively to individual rows
of cattle pens, effluent ponds and manure piles, and elevated methane mixing ratios were detected as far
as 25 km from the feedlot.
2.2.4. Use of Satellites in Locating Methane Emissions.
By measuring solar backscatter in the shortwave infrared, satellites offer rapid and powerful methods of
locating methane emissions (Jacob et al., 2016). This is potentially important in the land tropics where in situ
observations are typically few and infrequent. Although the utility of satellite observation has hitherto been
limited by low measurement frequency, retrieval accuracy, and vertical and pixel spatial resolution, newer
satellites, which include commercial satellites that offer pixel‐scene resolutions of the order of a few
kilometers and greater sensitivities to the tropospheric column.
Many earlier studies used the European MIPAS and SCIAMACHY instruments on ENVISAT (Bergamaschi
et al., 2013; Frankenberg et al., 2006; Houweling et al., 2014), which also retrieved CO2 in the 1.65 μm band,
and other IR‐active gases, with a coarse pixel resolution. More recently, the Japanese GOSAT (Hu et al.,
2018; Kuze et al., 2009) has provided global mapping of methane, albeit with averaging kernels in the mid-
troposphere and limited sensitivity to methane in the lower troposphere (e.g., Lange & Landgraf, 2018).
Although SCIAMACHY (2002–2012) and GOSAT (launched 2009) were capable of identifying regionally
important high column measurement over hotspots, the spatial and boundary layer resolution was poor
compared to in situ and aircraft mapping, and complicated by biases and poor accuracy of satellite measure-
ment. Moreover, it is not easy to obtain flux quantification from satellite observations. Thus, for the purposes
of reducing emissions, satellite observation to date has been of general but not specific value.
However, when used in the context of other inputs and chemical transport modeling, satellite results can be
helpful in targeting reduction measures. Specific local enhancements can be detected using GOSAT observa-
tions, as shown by Sheng, Jacob, Turner, et al. (2018), who used satellite results, together with a gridded

NISBET ET AL. 9 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

inventory of emissions (Maasakkers et al., 2016), to infer trends in North American emissions from different
source types (oil and gas, livestock, and wetlands). In an influential study, Kort et al. (2014) coupled satellite
observation with ground‐based validation to identify the largest satellite‐detected anomaly in the United
States within the Four Corners region in the Southwest United States (particularly coalbed extraction),
which they showed to have emissions up to 10% of the total U.S. methane emission inventory. In another
example, Wecht et al. (2014) used satellite and aircraft observations, coupled with the local inventory of
emissions, to constrain and attribute the geographical distribution of emissions in California. On an even
larger scale, Miller et al. (2019) used observations from the Japanese GOSAT satellite to assess China's
methane emissions.
Recent instruments are considerably advancing the power of satellite observation to help in locating emis-
sions, with better resolution and more frequent observation (Sheng, Jacob, Maasakkers, et al., 2018).
Turner et al. (2018) investigated the capabilities of different satellite observing systems. They found the
TROPOMI instrument, with 7 × 7‐km pixels, 11‐ppb single‐retrieval column‐weighted precision, and daily
frequency (Veefkind et al., 2012), should be capable of giving useful information on a spatial resolution of
about 30 km. The planned GeoCARB satellite, to be launched in the early 2020s, should give even better
resolution (2.7 × 3 km) and precision (4 ppb), making it potentially capable of usefully studying methane
emissions from large urban centers such as Shanghai (O'Brien et al., 2016), although aerosol pollution will
pose a major problem in retrieving accurate methane mixing ratios from regions like China, India, and tro-
pical Africa in biomass burning season. Cusworth et al. (2018) showed that TROPOMI and GeoCARB will be
successful in locating major sources of emissions in spread‐out gasfields but will be less successful where gas
installations are densely packed.
A powerful example of the usefulness of satellite detection was given by Pandey et al. (2019) who used
TROPOMI observations to discover a very large and hitherto little reported methane emission in the U.S.
state of Ohio in February–March 2018. The blowout continued for 20 days, and was observed by
TROPOMI on the thirteenth day, when emissions had likely declined from an initial peak rate. With an atmo-
spheric tracer transport simulation, Pandey et al. investigated the methane plume dispersion and used both
mass balance and cross‐section flux approaches to quantify the emission rate. They found an emission rate of
120 ± 32 metric tons per hour, which was twice that of the much better reported Aliso Canyon event (Conley
et al., 2016 and see below) and that the total methane emission in this little‐discussed Ohio event was perhaps
a quarter of Ohio's annual emission. This case demonstrates the power of satellite observation and its useful-
ness in tracking underreported events.
Finer detail is offered by instruments with pixel resolutions varying from 1–10 km down to as fine as 50 × 50‐
m resolution. MethaneSAT, with high spatial resolution and very high precision, is designed to monitor
emissions from about 50 oil and gas production regions accounting for about 80% of global production. It will
measure methane and carbon dioxide over a ~200km wide swathe (Wofsy & Hamburg, 2019).The high‐
resolution GHGSat microsatellites designed for greenhouse gas detection, observing methane columns over
selected 10 × 10‐km locations, give high effective pixel resolution of 50 × 50 m and 1–5% precision (Jacob
et al., 2016; Varon et al., 2018). Next‐generation versions with excellent spatial resolution and improvement
in point source detection better than ~10 kt/yr (0.01 Tg/yr) should be valuable where likely emission loci are
already known, for example, in identifying superemitters in gasfields, and in targeting locations of major
leaks on gas industry sites.
A different satellite perspective was taken by Elvidge et al. (2009, 2016) who used low‐light imaging data to
estimate global natural gas flaring. This major source of both CO2 and methane emissions remains poorly
quantified, especially in regions like Nigeria and Russia. Elvidge et al (2009) showed that in both these
regions the flaring efficiency markedly improved over the 2005‐2008 period. More recently, Deetz and
Vogel (2017) used visible infrared imagery to show continued decrease in Nigerian emissions of CO2 from
flaring. If it is assumed that as flaring reduced, methane emissions also declined, this methodology becomes
valuable in regional inventory construction.

2.3. Source Identification


Source identification depends on source flux, temperature, mixing layer height, wind speed, and wind direc-
tion. The problem is complex (Lowry et al., 2019). Often, major emissions from different source types are

NISBET ET AL. 10 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

closely juxtaposed. Facilities like waste handling and sewage treatment plants tend to be colocated in indus-
trial districts, often with a gas facility, a landfill, and perhaps an incinerator close by, surrounded in many
cities by dense suburbs. Sewage treatment plants can be located next to gas distribution installations, or
nearby landfills can emit gas close to waste burning facilities. In Australia, the United States, and Canada,
gasfields host major cattle feedlots, in some cases sharing water use and capable of emitting more methane
than the gas wells. In these settings, source attribution can be challenging. There are so many candidates for
methane emissions. Where multiply overlaid sources occur, help can come from of key identifiers such as
the presence or absence of associated species such as ethane, or isotopic signatures (Lowry et al., 2019).
Methane/ethane ratios are powerful discriminants between separate source signals, enabling identification
of specific emitters (e.g., Barkley et al., 2019; Feinberg et al., 2018; Kille et al., 2019; Mielke‐Maday et al.,
2019). Gas leaks contain significant amounts of ethane, while cows do not breathe out much. δ13CCH4 stu-
dies are powerful tools in pin‐pointing and segregating sources where urban land planning has produced
clusters of large but distinct point sources.
2.3.1. Identification by Methane/Ethane Ratio
In Los Angeles, Wennberg et al. (2012) observed the ambient air observations of methane, ethane, and car-
bon monoxide, in ambient air, and also measured ethane/methane enhancement ratios in the natural gas
supply. Assuming nearly all the ethane came from fugitive emissions from the natural gas supply, they were
able to attribute most of the excess methane measured in regional air to losses from the gas system, which
may have been losing 2.5–6% of its gas to the atmosphere. Similarly, Smith et al. (2015) used ethane‐to‐
methane correlations to quantify the fraction of CH4 emissions derived from fossil and microbial sources
in the Barnett Shale in Texas.
Using an aircraft, Peischl et al. (2015) measured multiple species in the planetary boundary layer over the
major U.S. Haynesville, Fayetteville, and Marcellus gasfields and showed the ratios between methane and
other alkanes in the regional atmospheric increment were the same as in local natural gas, thereby proving
the increment came from the gasfields. Also, in studies of five U.S. gasfields, Peischl et al. (2018) used
methane/ethane ratios to demonstrate emissions were from the gasfields and not from agricultural sources,
given similar atmospheric C2H6 to CH4 enhancement ratios and the composition of raw natural
gas withdrawn from the region. However, Lan et al. (2019) found an increasing trend in ethane/methane
emission ratios in U.S. sources, which may have led to significant overestimation of methane emissions
in some studies.
2.3.2. Isotopic Characterization
Isotopic characterization demands high‐precision measurement of the 12C:13C ratios (expressed as δ13CCH4)
of enhanced and background methane in ambient air masses (Keeling, 1958; Pataki et al., 2003; Zazzeri
et al., 2017). In settings where many sources can be geographically closely superposed, isotopic signatures
are powerful discriminators between sources. To take an extreme example, an aircraft survey found a
methane‐rich plume downwind of the U.K.'s East Anglian offshore gasfields, but isotopic study showed that
a significant part of the methane anomaly was not from the gas platforms but had blown there from remote
on‐land sources (Cain et al., 2017). They found a significant methane anomaly in a plume between 25 and 50
km wide, with elevated methane at 70 to 100 ppb above the background, which was detected during a low
altitude (100 masl) overflight over the Leman field's gas platforms in the southern North Sea, off the coast
of Norfolk, England. The methane plume was in close proximity above large gas platforms. Air samples were
taken and surprisingly showed much of the methane anomaly had an isotopic signature δ13CCH4 of −55‰, a
characteristic ratio for land sources such as cattle or landfills, and not the −32‰ signature of these gasfields,
showing that much of the source was not the gas platforms.
In ambient air relatively remote from sources, off‐line lab‐based isotope ratio mass spectrometry is required
to reach the needed analytical precision, by analyzing air samples taken at the study sites. This is because
typically it is necessary to determine δ13CCH4 to a precision of 0.05‰ or better (Fisher et al., 2006) in ambient
air masses. Adequate precision is currently achievable by vehicle‐mounted traveling optical analyzers only
when very close to source in methane‐enriched air. These instruments typically have isotopic precisions
of around 2–3‰, though improved precision can be attained, for example, if the plume can be sampled by
AirCore (Karion et al., 2010; Rella et al., 2015) for isotopic playback. An alternative approach (Lu et al.,
2019) that shows promise is to use a moving Miller‐Tans analysis (Miller & Tans, 2003) of time series data
to characterize the population distribution of the isotopic signature for a given source.

NISBET ET AL. 11 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Townsend‐Small et al. (2012) used isotopic ratios to interpret inputs to air samples collected across the Los
Angeles megacity. They used vehicles to access a variety of sources including oil well fields, oil refineries,
power plants, traffic, cattle and cattle feedlots, landfills, and sewage facilities. In addition, they used the
geography of the region to collect air from a local mountain (1,707 m above sea level). Methane in the
ambient air samples was then analyzed for δ13CCH4 and δDCH4 as well as for Δ14CCH4. They found that in
Los Angeles the main source of methane emission was leakage of gas from fossil fuel use, including
geological sources, natural gas pipelines, oil refining, and power plants. In the Los Angeles study, δDCH4
was the strongest indicator of methane source.
This result from Los Angeles was in contrast to the older results of Lowry et al. (2001), also using stable
isotopes, who found a major source of methane emission in the London area was landfills. Lowry et al.
(2001) were reporting on London around the millennium: More recently, much better landfill practice
has led to lower U.K. landfill emissions (UK NIR, 2019). Currently, Royal Holloway laboratory
observations suggest leaking natural gas is the largest contributor to methane emissions around Egham,
in western London. This is because all the landfills west of London in the 2001 study have now closed,
and gas leaks now predominate.

Zazzeri et al. (2015) used Keeling plots to characterize sources, using bag samples. Figure 6a is from a landfill
at Mucking in SE England, emitting methane with δ13CCH4‐56.1 ± 0.5‰ (2 SD) while the Keeling plot in
Figure 6b is from a gas works emitting methane with δ13CCH4‐36.3 ± 0.3‰. The results show that these
two sources are readily distinguishable. In contrast, Figure 6c shows results from three coal mines. The
δ13CCH4 intercept from the Thoresby mine in Nottinghamshire, England, was −51.2 ± 0.3‰, markedly dif-
ferent from the methane emitted by the Aberpergwym and Unity mines, which were neighbors 1.2 km apart
in the South Wales coalfield. Emissions from the latter two were isotopically similar, with signatures around
δ13CCH4‐32‰ that are just within error and very different from the first mine. Intrinsically methane from
these two coalmine emission sources may be separable, with a high enough sample population and also
using careful directional measurement by a mobile system. Thus, these results show that even with closely
similar sources, such as neighboring coal mines or gaswells, it may be routinely possible to identify emis-
sions by off‐site surface measurement from public access roads. A similar methodology has been used for
CO2 by Domenikos et al. (2019), while Fries et al. (2018) used carbon (δ13CCH4) and hydrogen (δ2HCH4)
stable isotopic determination to distinguish between biogenic methane from sewer gas and thermogenic
methane from leaking natural gas pipelines in Cincinnati, Ohio.

In these studies, for each emission source type, the isotopic signature and/or ethane:methane ratio must be
known for the specific location being investigated. However, both isotopic signatures and ethane:methane
ratios vary widely depending on the source, whether gas basin or farm location, and in some cases significant
variation in these signatures can occur very locally.

Case study A: attribution of sources around Fylde, UK. Source attribution by isotopic measurements was used
to identify the emissions measured around Fylde, Lancashire, northern England, UK (Figures 7 and 8). The
measurement, reported here, was carried out as a baseline survey around a proposed site for gas extraction by
fracking. Gas sources are identified by their methane/ethane ratios (Figure 9) and their different isotopic sig-
natures. Farm signatures have δ13CCH4 close to −60‰, manure piles −50‰, and gas leaks −40‰. In a similar
study in the Vale of Pickering, also in northern England, gas leaks from the natural gas offtake station had
δ13CCH4 of −42‰, while farm cow manure piles had δ13CCH4 of around −50‰, and methane from the cattle
in the barns themselves and their waste had a combined δ13CCH4‐59‰. However, methane from the local
landfill had δ13CCH4 of −57‰, closely similar to methane from cattle eating mixed C3/C4 fodder (e.g., grass
plus maize). Thus, wind direction is needed to discriminate between landfill emissions and cattle breath.

Case Study B: unexpected events: Elgin gas leak. Aircraft are important when major unexpected events occur.
An uncontrolled gas release took place from the Elgin platform in the U.K. North Sea from 25 March to 16
May 2012 (Lee, Mobbs, et al., 2018). When the gas blowout occurred, the U.K. FAAM (Facility for
Airborne Atmospheric Measurement) BAe146 aircraft was fortuitously fully equipped as it had just returned
from measuring methane in Scandinavia: It was immediately sent to fly over the gas platform (Figure 9).
Flying at low altitude, FAAM was able to map out and sample the plume of methane‐rich gas from the leak
(Lee, Mobbs, et al., 2018). This study was carried out flying normal to wind direction. As a general rule

NISBET ET AL. 12 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 6. Keeling plots from (a) gas works in Staines (UK), (b) landfill (UK), and (c) three coal mines in the United Kingdom. Errors on the y axis are within 0.05 ‰
−t
and on the x axis within 0.0001 ppm and are not noticeable on the graph. From Zazzeri et al. (2015).

however, where the source is known, it is better to fly at an angle (e.g., 30°) to wind direction in order to
maximize the intersection.
Isotopic measurements (Figure 9) then helped pinpoint the source of the gas in the blowout. Keeling plots in
air samples from flights transecting the gas plume (Figure 9) found the source gas had δ13CCH4 of ‐42.3 ±
0.7‰. The results pointed to a source in a small gas pocket, geologically shallower than the main gasfield
and that the potential volume of gas involved in the leak was limited. The hypothesis was borne out. The leak
diminished rapidly. This FAAM aircraft study informed the platform operators, advising them that the gas

Figure 7. Fine detail of road circuit around a proposed U.K. shale gas drilling site (Fylde, Lancashire, UK) showing the main sources and their isotopic signatures,
13
and the excess methane averaged across 18 days of surveys and aggregated into 10 m × 10 m grid squares. Farm δ CCH4 signatures are close to −60‰,
manure piles −50‰, and gas leaks −40‰. RHUL results.

NISBET ET AL. 13 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 8. Distinct methane/ethane ratios for different sources, Fylde study, RHUL results.

flux from leak was reduced. The leak was then ended by accessing the platform for a top kill, rather than
drilling down for a much slower bottom kill that had been discussed.
A much larger event followed 3 yr after Elgin, when a blowout took place from the Aliso Canyon under-
ground storage facility in California in 2015. This emission was studied by a team which included partici-
pants both from the effort to quantify the Elgin leak and also from the study of fracking in U.S. gasfields
(Conley et al., 2016). In an impressive campaign, they flew dozens of plume transects over the region, as well
as taking whole air samples on the ground downwind. This successfully mapped out the dispersion of the
leaked gas. However, in contrast to the Elgin study, which used δ13CCH4 in an attempt to identify the source
reservoir, Conley et al. (2016) measured ethane, C2H6, as the marker to distinguish the emissions from the
natural gas leak from other sources of methane.

3. Quantification of Emissions
If a big leak occurs, it needs to be fixed. Rapid assessment that a leak is big is essential for safety. Unsafe leaks
must be stopped immediately. Thus, in many cases, the only necessary regulatory quantification is the dis-
tinction between “large” or “small.” For mitigation however, flux quantification is important. The most
obvious challenge is to find those neglected leaks that are large enough to be a significant emission, yet
below the “unsafe” or “must‐be‐fixed” threshold. Smaller leaks are of course also important for mitigation,
but how should mitigation efforts be targeted? To set priorities for cost‐effective reduction, leak quantifica-
tion is needed.
3.1. Direct Measurement
The simplest way to quantify an emission is to enclose it and measure it directly. Kang et al. (2014) and Lamb
et al. (2015) placed enclosure chambers around gas leaks and then directly sampled the methane that was
captured. Using this methodology, Kang et al. (2016) demonstrated that 5–8% of annual methane emissions
in the U.S. state of Pennsylvania came from abandoned oil and gas wells, many of which were high emitters
of methane. Many of the highest emitters are in the Appalachian basin, including Pennsylvania, Ohio, and
West Virginia, and they are among the oldest oil and gas wells in the world, some over 100 or 150 yr old
(Townsend‐Small et al., 2016).
Abandoned wells are relatively straightforward (though not necessarily easy) to locate and once identified
may be quickly quantifiable within broad error limits. Although unlined abandoned holes may be difficult

NISBET ET AL. 14 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 9. (left) Elgin gas production platform, from the window of the U.K. FAAM aircraft, taken during sampling on 2 April 2012. Photo: Mathias
Lanoisellé (RHUL group). (lower right) 30 March onboard CH4 plot for measurements 5 nautical miles (9.26 km) downwind of the leak; peaks are for measure-
ments at different altitudes—from left 37, 153, and 233 m (see Lee, Mobbs, et al., 2018). (upper right) Keeling plot of air samples from Elgin methane plume.
13 13
Inset—enlarged detail. δ CCH4‐42.3 ± 0.7 ‰. (2σ error: geometric mean regression). Measurements of CH4 and δ CCH4 at RHUL (Lee, Mobbs, et al., 2018).

to plug, given the cost of both the leak and the explosion danger such super‐emitters pose, they should
usually be an obvious target for mitigation. However, many of these historic wells are “orphans” that do
not have any permit or current owners, leaving the responsibility for plugging to the state or federal
government, which is slow, and mitigation prioritization may not be based on methane emission rate
(Townsend‐Small et al., 2016).

3.2. Mass Balance and Tracer Quantification Methods


For most sources, fluxes need to be assessed by measuring the air into which the emissions are mixing. In
most circumstances it is not possible to enclose sources in chambers, because the sources are unknown, phy-
sically too large or disseminated, or otherwise uncapturable. That poses the difficult problem: How does one
quantify emissions that are input into moving air masses? The problem is exacerbated, especially in less
developed nations, by the lack of access or skills to use sophisticated meteorological products. What may
be possible on a flat North American prairie with funding for real‐time local meteorology, good computing
skill, and access to sophisticated modeling software may not be feasible for a low‐income community trying
to assess fluxes from a landfill located among complex topography under the intertropical convergence zone.
One of the simplest mass balance quantification methods was used by Lowry et al. (2001), who studied the
London conurbation, which has a detailed CO2 emissions inventory. They measured the excess CO2 and
CH4 inputs to urban air, by subtracting the known mixing ratios in contemporary Atlantic background
air, and then obtained the methane emissions as a ratio to the inventory CO2 emissions, assuming the
bottom‐up CO2 inventory to be valid. This allowed them to test the less well constrained London CH4 emis-
sions inventory for consistency against the better‐constrained CO2 emissions inventory.

NISBET ET AL. 15 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Proxy tracer (e.g., CO, N2O or acetylene, or additional CH4) release methods, using deliberate emission of a
tracer gas at a known mass/second rate to validate a dispersion model also offer important techniques to
quantify fluxes (Bell et al., 2017). The method relies on the assumption that the released tracer gas will
behave in the same way as methane emitting from the landfill. The methane emission is then proportional
to the ratio of the integrated concentration of the emitted methane to the integrated concentration of the
released tracer gas, as measured across the dispersion plume (Scheutz & Kjeldsen, 2019). Tracer release
can be coupled with spectroscopy to calibrate fluxes from individual sites such as landfill and oil and gas
facilities (Ars et al., 2017; Foster‐Wittig et al., 2015; Rees‐White et al., 2018; Roscioli et al., 2015; Scheutz
et al., 2011). Indeed, Scheutz and Kjeldsen (2019) suggest that the tracer gas dispersion measuring technique
should be a core methodology in monitoring to determine emissions.
The aircraft mass balance approach was used by Ren et al. (2019) to estimate emissions from unconventional
gas and oil extraction from the Marcellus Shale in Pennsylvania and Virginia, over a 4,200‐km2 area. They
found emission rates around 0.8% to 1.5% of total gas production, lower than some other studies but
consistent with the national inventory assessment. In another study, over major U.S. East Coast cities,
Plant et al. (2019) used aircraft observations of CH4/CO2 ratios and other species to test local emissions
inventories. They found that emissions, predominantly attributed to fugitive natural gas, were very
substantial and more than twice the recent gridded inventory estimates. Wunch et al. (2016) used a similar
technique with CH4/C2H6 ratios to estimate methane emissions in Southern California, another large
heavily urbanized urban area. Using historic data, Wunch et al. (2016) analyzed records of total column
abundances for ethane and methane and studied the historic evolution of ethane to methane ratios. They
showed that more than half of the excess methane in the Southern California coast basin's air came from
natural gas losses.
Pitt et al. (2019) quantified fluxes using a mass balance approach for the Greater London area using the U.K.
FAAM research aircraft (www.faam.ac.uk) equipped with a cavity‐enhanced absorption spectrometer. The
same aircraft was used by Lee, Mobbs, et al. (2018), to quantify emission flux from an uncontrolled methane
leak from an oil platform in the North Sea. Likewise, O'Shea et al. (2013) used a cavity‐enhanced absorption
spectrometer for airborne measurements over Arctic Scandinavia. With simple box modeling and boundary
layer measurement to obtain methane fluxes from wetlands, they found results consistent with parallel in
situ chamber‐based measurements on the ground below (O'Shea, Allen, Gallagher et al., 2014). With a dif-
ferent approach in flights over England, Pitt et al. (2016) used an airborne quantum cascade laser absorption
spectrometer to measure methane and N2O, demonstrating the effectiveness of the method. Moreover,
aircraft‐based infrared remote sensing of methane (and other proxy tracers) can offer higher spatial resolu-
tion retrieval compared to satellites (e.g., Allen et al., 2014).
Airborne remote sensing with a visible/infrared imaging spectrometer is extremely powerful providing a
regional assessment. In the California Methane Survey, Duren et al. (2019) measured ground‐reflected solar
radiation between 380 and 2,510 nm, from an aircraft at 3‐km altitude, over oil and gas fields, as well as areas
with manure and waste management activities. In their campaigns they covered 272,000 facilities, including
over 200,000 oil and gas wells. They detected 564 strong point sources emitting methane, especially from
landfills. For these distinct sources, measured at 250 facilities, they observed the local methane enhance-
ment and then estimated emission fluxes using wind speed data from weather reanalysis products. They
found a population with a heavy tail distribution, with 10% of the point sources responsible for 60% of the
emissions. In particular, the largest category of super‐emitters was landfills, especially a subset of superemit-
ters, while dairy farms and oil and gas facilities also contributed.

3.3. Mathematical Modeling


An effective and simple quantification method uses the assumption that the source is a point and that the
plume disperses downwind (Brandt et al., 2014; Robertson et al., 2017; White et al., 1976). When methane
is emitted from a source, the input is entrained into the moving mass of ambient air. In this moving air mass,
the input then disperses both horizontally and vertically, to form a cone of dispersion. If it is valid to assume
the gas is well‐mixed in this cone, then the mole fraction of the gas at any nearby point where it can be mea-
sured depends on the source strength and height, the wind speed, and the stability of the air mass (Seinfeld &
Pandis, 2006; Pasquill, 1975).

NISBET ET AL. 16 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Caulton et al. (2018) very successfully used a Gaussian methodology and a hierarchical measurement
strategy involving both mobile vehicle‐mounted measurement and small (2–3 m) deployable towers, as well
as controlled releases, to quantify emissions from a field of over 1,000 gas wells in Pennsylvania, USA. They
clearly demonstrated it is possible reliably to identify and roughly quantify extreme emission sites (i.e.,
superemitter mitigation targets).
Modeling tools are accessible via U.S. Environmental Protection Agency (EPA) “Other test method”
OTM 33a (via https://www3.epa.gov/ttnemc01/prelim/otm33a.pdf). This is a way of identifying and
roughly quantifying emissions from sources located by large spikes in the analyzer record as the vehicle
drives past; for example, it is ideal for low‐cost regulatory measurement of emissions from gas wells and
pipe leaks.
Eddy covariance techniques have long been used to quantify emissions in natural settings (e.g., Detto et al.,
2011; Rinne et al., 2007). In Florence, Italy, Gioli et al. (2012) sampled from a 3‐m mast that was on a roof
14 m above average roof level and 33 m above ground. Methane fluxes from the footprint area were
189.2 ± 7.0 mg CH4·m−2·day−1 and did not show significant temporal variability. Similarly, Helfter et al.
(2016) measured CH4, CO2, and CO for 3 yr from the BT Tower in central London, 190 m above street level,
and from another rooftop site 2 km away, 50 m above street level. Methane emissions in London were
similar to Florence, at 197 ± 8 mg CH4·m−2·day−1, but were found to be double inventory assessments
and showed moderate seasonality (21% larger in winter). These studies demonstrate the power of the eddy
covariance technique in urban settings and its usefulness in testing bottom‐up inventory assessments. In
principle, future drones fitted with better sensors and good anemometers will carry out eddy flux
calculations directly.

3.4. Modeling Use of Tall Tower Networks


In densely populated urban areas, or in major producing regions with dense sources, such as many onshore
gasfields, well‐located measurement towers can be used to pinpoint emissions and to assess regional emis-
sions using eddy covariance or optimized Lagrangian‐inversion techniques. This can be applied on any scale,
from city to continent or planet. For the US Los Angeles Megacity, Yadav et al. (2019) took data from a net-
work of 15 in situ sites and used inverse modeling to estimate methane emissions from 3‐km cells, over 4‐day
windows, across the megacity area. This method located emissions in enough detail to pinpoint major
sources, for example correctly detecting the shutdown of emissions from a major landfill. However, for
the very large and highly localized Aliso Canyon leak there were periods when the inversions did not capture
fluxes well, especially in unhelpful prevailing wind conditions.
More widely in the United States, Lan et al. (2019) used long‐term data from both aircraft and tall
towers to analyze methane emissions from a variety of sites around the country, including in oil and
gas production areas. Encouragingly, this important study found that in the period 2006–2015 the
increase in emissions from North American oil and gas industry sources was much smaller than find-
ings in many previous estimates.
In Europe, Bergamaschi et al. (2018) used data from measurement networks to validate regional emissions,
and Wunch et al. (2019) used long‐term, ground‐based measurements of atmospheric total column methane
abundance by remote sensing, to assess methane emissions from the wide northern European region from
Poland to France. Their results implied that the European inventories were likely overestimated. The
Integrated Carbon Observing System network maintains a long‐term network of stations carrying out
high‐precision greenhouse gas measurement (https://www.icos‐ri.eu/icos‐stations‐network). For the
United Kingdom the GAUGE (Greenhouse gAs Uk and Global Emissions) Project used a mix of tall tower
(Stanley et al., 2018; Stavert et al., 2018) and other ground‐based measurements, supported by other air‐
borne and ship‐borne measurements and satellites with the aim of producing better understanding of
U.K. emissions (Palmer et al., 2018).
While tall tower networks are appropriate as designed for regional budget estimates, their continuous
records of methane mole fraction can intrinsically also be used by back‐trajectory analysis to locate major
nearby sources, especially if supplemented by air‐sampling for high‐precision isotopic or alkane ratio mea-
surement to differentiate between source types such as landfills or gas leaks. Henne et al. (2016) validated
the Swiss methane inventory using observations from a network of sampling sites in the Swiss plateau and

NISBET ET AL. 17 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

also in the Alps, as the basis of inverse modelling with high spatial resolution <10km over Switzerland. In
Australia, Luhur et al. (2018) used two towers in conjunction with inverse modeling to assess regional
methane emissions and quantify fluxes in the Surat Basin from coal seam gas production and processing,
coal mining, cattle feedlots, piggeries, wetlands, and other anthropogenic activities. The model domain was
350 × 350 km, which demonstrates the cost‐effectiveness of towers for making a top‐down estimate check
of bottom‐up inventories.

3.5. Practical Methodology


In practice, each quantification problem is different, and measuring teams have different analytical assets.
Ground vehicles are appropriate and widely used for studying locally sourced plumes (e.g., Baillie et al.,
2019) and can be supplemented with UAV (drone) grab‐bag sampling (Brownlow et al., 2016), or by tethered
sampling tubes lifted either by UAVs (Allen et al., 2019), or pop‐up balloons (Steiger et al., 2015). Such
approaches can yield 3‐D sampling for local‐scale (e.g., site‐specific) flux retrieval using mass balancing
approaches to quantify fluxes accurately. In the cases where a measurement team has a mobile analyzer
in a vehicle with a sonic anemometer and GPS location (as in Figure 3, left), and where low‐level gas plumes
can be measured within a few hundred meters of the source, methane emissions can be calculated by driving
through the plume. If the source of the emission is known, tracer release can also be an accurate method for
assessing methane emissions (Lamb et al., 1995, 2015). Isotopes are particularly helpful in apportioning
emissions where various sources are present (Townsend‐Small et al., 2015).

As well as cars, bicycles, and child buggies are useful for carrying instruments and batteries, as are
backpacks. In a detailed study of emissions around the Munich Oktoberfest in Germany (a major beer
festival), Chen et al. (2019) made in situ measurements by walking and biking around the perimeter of
the Oktoberfest. They then applied a Gaussian plume dispersion model to assess emissions.
Aircraft are appropriate in assessing emissions from large conurbations and regions, to verify inventories
from wider areas, where there are many sources, such as large gasfields or cities, or wide zones of natural
emissions (Miller et al., 2016). For example, Cui et al. (2019) quantified emissions from the Haynesville‐
Bossier oil and gasfield by using inverse model calculations driven by aircraft measurements over the gas-
field, and showed that emissions were probably larger than estimated in the national inventory. Over the
U.S. city of Indianapolis, Cambaliza et al. (2015) and Heimburger et al. (2017) used a light aircraft and a
mass‐balance approach as well as methane:propane ratios in air samples to quantify urban emissions.
O'Shea, Allen, Fleming, et al. (2014) used upwind and downwind airborne measurements and a mass bud-
get box modeling approach to determine net regional flux of methane for Greater London. In an important
study of the Fayetteville shale play, a major U.S. gas producing region, Schwietzke, Pétron et al. (2017)
used aircraft measurement and a mass balance approach, integrating methane mole fraction measure-
ments across plume widths in known wind velocities and direction, to assess methane emissions. The
methodology of O'Shea et al. (2013), O'Shea, Allen, Fleming, et al., 2014. O'Shea, Allen, Gallagher, et al.,
2014) and Hartery et al. (2018), though designed to locate and measure natural emissions, is well suited
as a basis for verification of emissions inventories on a regional scale.
Frequent repeated surveys by light aircraft can track seasonal changes in emission patterns and inform local
regulatory bodies of unexpected new inputs or successes in reduction policies. Though derived fluxes can be
precise but highly inaccurate, when coupled with technological developments in sampling platforms and
new advances in flux inversion modeling such as mass balancing, plume inversion, and gradient flux tech-
niques, calculated fluxes can be used to identify and quantify individual sources and source types (Karion
et al., 2019).
Schwietzke et al. (2018) and Englander et al. (2018) showed that in the U.S. context aerial surveys on light
aircraft are an excellent tool, able to locate anomalously highly emitting sources and thus able to target on‐
the‐ground vehicle‐based identification and inspection. It is thus becoming feasible to identify leaks where
fixing the emission is most cost‐effective (Schwietzke et al., 2018). In some cases, it is likely the combination
of aerial overpass and ground follow‐up to mitigate emissions can be cost saving and thus profitable to the
industry, in other cases, mitigation may have to be driven by regulatory pressure. Aircraft observation of sus-
tained major methane emissions stimulates efforts to reduce unprofitable leakage.

NISBET ET AL. 18 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

The usefulness of aircraft in major events (see Case Studies A and B above) lends weight to the case for
national bodies to maintain aircraft capability as a standing asset for dealing with major air pollution events,
also useful in assessing dust clouds from volcanic eruptions, etc.

These intuitive points can be made:


1. For many companies, better aerial and ground identification of leaks should now become a priority as a
source of profit and improved safety, in addition to greenhouse gas reduction.
2. For regulatory bodies, the costs of leak detection are declining and leak surveys are becoming more
affordable as instruments improve. Leak detection is thus now becoming more affordable, especially
when there is potential income from penalties and fines.
3. Methane surveys are likely also to detect sources of emissions that are not being captured in national
inventories.
4. Where sudden accidental major leaks occur, urgent measurement is both essential and powerful in
rapidly assessing the event and in helping to end it. Thus, wise policy is to maintain ongoing deployable
capability for such events, including aircraft, at the national level.
5. There are persuasive grounds for regulatory jurisdictions to maintain aircraft capability for leaks as well
as other air pollution events.
6. Hitherto, Drone/UAV measurements have lacked adequate precision to support accurate emission quan-
tification. This is changing rapidly, for example, with new midinfrared lasers offering better sensor pre-
cision and/or lighter analyzers. However, it should be noted that with increasing regulation of UAV
flying, especially near facilities such as gas installations it may become preferable to use pop‐up balloons,
for example, for sampling.
7. On a national or regional scale, various methodologies are available to quantify emissions, but many of
these need access to prior inventories, whether CO2 or CH4, together with relatively sophisticated
facilities (such as aircraft) and advanced modeling techniques. Such methodologies are useful in
developed nations, but in many tropical nations may not be accessible. Here upwind/downwind simple
mass balance techniques and single analyzer measurements (CO2 and CH4) may suffice. There is much
need for better quantification methodologies that are more accessible to scientists and regulators in
less‐developed nations.

4. Practical Emission Reduction and Removal—Tractable emissions


“Tractable” emissions can be defined as those from easily located point sources, for which, once the points
are found, mitigation measures can readily be undertaken. In most heavily populated regions the methane
source mix is diverse (e.g., Lowry et al., 2001) but widely tractable to mitigation. Such emissions come from
many sites—the gas industry, urban wastewater and sewage, landfills, etc. The development of low‐cost
methodological advances in leak detection, to identify superemitter leaks, should have a significant impact,
especially if better UAV systems become widespread in the near future. It is bad management that leaks gas.
It is much in corporate interest to eliminate super‐emitters as they likely cause both loss of profit and also
potentially expensive safety risks.
Obvious examples of tractable emissions are known deliberately vented emissions in gas fields and also
many emissions in urban distribution systems, both deliberate vents and known leaks that are not large
enough to be classed as safety hazards and hence neglected. At the other end of the tractable spectrum
are leaks from abandoned installations, and small harder‐to‐find leaks from urban systems. In these cases,
deliberate policy is needed to mitigate emissions. However, this is not necessarily very expensive. Kang
et al. (2019) found that reducing methane emissions from abandoned oil and gas wells has comparable costs
to other greenhouse gas mitigation options.
Typical examples in gasfields (Vaughn et al., 2017; Vaughn et al., 2018) would include methane emitted not
only from readily discovered leaks around gaswells but also from deliberate operations such as flaring, vent-
ing, and pipe maintenance; unburned gas entrained in compressor exhausts during operation; and also emis-
sions from known processes such as water handling, which can be redesigned or better contained. Episodic
venting (Vaughn et al., 2018), which is deliberate, is particularly addressable. Leaks from urban gas

NISBET ET AL. 19 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

networks, landfills, and sewage treatment facility emissions, are widely tractable, as they are known, easily
located, and predictable, as are emissions from underground coal mines, where shaft and ventilation outlet
locations are well known.
In large gas production and urban distribution systems, discovery and quick approximate quantification is
arguably nine tenths of mitigation: Once large sources are known, emissions can generally be ended quickly,
often cheaply, even profitably, and with benefit to safety. Both deliberate venting and accidental fugitive
releases are readily reducible. Linked to such mitigation measures, strong regulatory efforts can reduce vent-
ing and flaring (for U.S. data see https://www.eia.gov/dnav/ng/ng_prod_sum_a_EPG0_VGV_mmcf_a.
htm). Flaring and venting gas not only does environmental damage; it also reduces the long‐term energy
security of the nation that tolerates the flaring.

4.1. Gasfield Superemitters


To cut emissions the most obviously tractable targets are the so‐called “super‐emitters” from the wells, pipe-
lines, pumping stations, and urban distribution networks of natural gas systems. Brandt et al. (2014—see
especially their Table S6) showed that in many U.S. gasfields and gas transmission and distribution systems,
much of the mass of gas emitted comes from a few major leaks. Similarly, a recent synthesis of methane
emissions from the oil and natural gas supply chain indicates that the most emissions come from production
wells and that those are dominated by a few major sources (Alvarez et al., 2018). Mineral deposits provide a
good analogy. They typically follow Zipf's law of distributions (Guj et al., 2011): in gold mining, there are
many small mineral deposits, but few of largest rank, and most of the extractable gold comes from a few very
large gold mines. Intuitively, gas leaks would be expected to show similar power law patterns, with a few
large leaks and many small, and with a significant proportion of the total amount of leaked gas coming from
the few large sources.
Detection and monitoring of super‐emitters is rapidly becoming simpler with rapid advances in vehicle‐
based sensors (Jackson et al., 2014; Zazzeri et al., 2015), aircraft measurement (Schwietzke, Pétron, et al.,
2017) and satellite sensing (Jacob et al., 2016). Satellite observations can now attain high spatial resolution
(50 m) at known sites, and are excellent for discovering large emission point sources in remote locations
(Varon et al., 2019), but currently remain too sparse to constrain emissions by mapping extensive regions
at fine local detail (Turner et al., 2018; Wecht et al., 2014). However, the detection and quantification
of a major leakage event in Ohio by the TROPOMI instrument (Pandey et al., 2019; see section 2.2.4)
demonstrates that satellite observation is rapidly becoming more able to find and quantify big leaks
from superemitters.
Measurement by unmanned aerial vehicles (e.g., see the GHGmap project, http://www.geosciencebc.com/
projects/2016‐065/), is improving with the development of inexpensive, low‐weight, yet more precise detec-
tors. However, restrictions on UAV flying near “sensitive” facilities like gas installations, may impose signif-
icant limits on the future usefulness of UAV‐mounted equipment to monitor facilities, detect leaks and
hence mitigate emissions.
Aircraft, UAV, and vehicle surveys can now be augmented by direct on‐site continuous monitoring, which is
now feasible around production facilities, to watch superemitters and potential super‐emitters across their
life cycle, using low cost sensors such as those described by Collier‐Oxandale et al. (2018). Similarly, to
achieve higher precision on large offshore gas platforms, automated cavity‐based analyzers can be linked
to air inlets distributed around the platform (e.g., an “octopus” design, with air hoses feeding via a central
multivalve to the analyzer).
In the Barnett Shale gasfield in Texas, superemitting sites accounted for roughly three fourths of total emis-
sions (Zavala‐Araiza, Lyon, Alvarez, Palacios, et al., 2015), which were 90% higher than inventory estimates
and constituted 1.5% of natural gas production (Zavala‐Araiza, Lyon, Alvarez, Davis, et al., 2015). Similarly,
though with local variation, in four other major U.S. gasfields the bulk of emissions came from 20% of sites
(Robertson et al., 2017). Gathering facilities are particularly important sources, especially the leakier instal-
lations: in the United States, Mitchell et al. (2015) found 30% of gathering facilities contribute 80% of the total
emissions. Abandoned wells also contribute significantly in some regions (Kang et al., 2016; Townsend‐
Small et al., 2016). Similar patterns of superemission from abandoned wells and coal‐mine vents were also
found in Queensland and New South Wales, Australia (Day et al., 2015; Zazzeri et al., 2016). It will likely

NISBET ET AL. 20 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

need public intervention to locate and cap the multitude of leaking abandoned exploration holes in gas and
coal extraction fields. This will at times demand deliberate significant public expenditure for old holes that
are “orphans,” with no identifiable or solvent corporate owner.
Measurements by Pétron et al. (2012) and Karion et al. (2013) showed very high leakage rates in U.S. shale
gasfields: In particular, in the Uintah gasfield in the western USA, Karion et al. (2013) showed emissions
were as high as 6.2% to 11.7% of average gas production. These studies have been heavily cited in academic
journals and were particularly effective in drawing attention to high loss rates. However, recent gasfield
studies elsewhere (e.g., Schwietzke, Pétron, et al., 2017—see above) have found much lower leakage rates,
typically well below 2%.
Leaks are not the only major cause of gasfield emissions. Schwietzke, Pétron, et al. (2017) and Vaughn et al.
(2017) showed that in the Fayetteville Shale, in Arkansas, USA, deliberate manual actions drove major after-
noon emission peaks. The western half of the gasfield had much higher emissions (1.8% of production) than
the eastern part (0.8%): This was because the water content of the gas was higher in the west, and the higher
but episodic emissions could be traced to manual unloading of accumulated water and other liquids. The
liquid unloading was by deliberate action for water management.
Deliberate venting of unwanted gas is a major source of emission in many oilfields. O'Connell et al. (2019)
studied heavy oil wells around Lloydminster, Canada. Over 40% of sampled well pads were emitting detect-
able methane, and of these, 40% emitted above the venting threshold beyond which mitigation was required
by Canadian federal regulations. Similarly, in the Surat Basin, Australia, high‐point vents on the coproduced
water distribution lines were found to be important sources of emissions (Day et al., 2015). Major sources of
methane emissions from venting and water handling (e.g., Figure 10; Iverach et al., 2015) should be feasible
to control, as it should be easy for even ill‐equipped local government authorities to detect and thus regulate
them. Good design is generally better than post‐installation mitigation. Thus, when developing new gas-
fields, consideration should also be given to reducing the number of high‐point vents on the coproduced
water distribution pipelines, and where such venting points are required the emissions from these points
should be captured.
Development of effective low‐cost leak detection and monitoring, coupled with targeting superemitters and
working with operators to improve operational practices, particularly in water handling, can enable rapid,
low‐cost cuts in emissions (Mayfield et al., 2017). Mitchell et al. (2015) pointed to the benefits from simple
vigilance: Gas processing plants were staffed and had comparatively low leakage; gathering facilities, typi-
cally unstaffed, had larger leaks. In major gasfield regions, even poorly resourced regulatory bodies should
be able to afford vehicle‐mounted detection systems of the type shown in Figure 3, which, by imposition
of large fines for detected superemitters would likely pay for themselves.
Regulation, assisted by tax and penalty nudges incentives, is an obvious first step to reduce leaks and flaring
from gasfield sources. In considering mitigation and safety (and perhaps profit), self‐interest can be made to
converge with greenhouse good. In some superemitter cases the economics of mitigating very large emis-
sions are so attractive that high leakage simply denotes bad cost control and incompetent corporate leader-
ship. If overall profit margins are, for example, 10%, a small investment to recapture a 10% leak will double
profits, and enhance safety. Canada has introduced some controls, intending to cut methane emissions by
40–45% by 2025, but clearly ineffective as yet in the case of the Lloydminster emissions reported by
O'Connell et al. (2019). Tyner and Johnson (2018), studying oil production sites in Alberta, Canada, used
Monte Carlo simulations to show that a 45% reduction in methane emissions from flaring and venting
would, per ton of CO2 equivalent greenhouse warming, cost between about CAN $2.5 to CAN $−3 (i.e., a
profit). Relatively small government tax incentives for mitigation and relatively large penalty disincentives
imposed on leakage would likely serve to nudge the mitigation of emissions well beyond a 45% cut. In any
nation, even such relatively lax requirements do help cut emissions, with the side impact of slightly enhan-
cing national energy longevity and security of supply.
In 2016, as part of the United States' plan to meet its Paris Climate goals, the U.S. EPA introduced the Oil and
Natural Gas Sector New Source Performance Standards, with a variety of emissions reductions requirements
wells and compressors constructed after September 2015, including regular leak detection and repair for
wells and compressor stations. However, EPA has challenged this rule in court and has reduced the amount
of monitoring required since the regulation was implemented. The U.S. Bureau of Land Management also

NISBET ET AL. 21 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 10. (a and b) Methane mole fraction plumes mapped during a mobile survey undertaken by UNSW and RHUL in
the Surat Basin in 2014 and presented here. There are 22 similar water holding ponds throughout the Surat Basin.
These ponds are required for managing the hundreds of megaliters of daily co‐produced water associated with coal seam
gas (CSG) production. Map data: Google DigitalGlobe.

implemented a Waste Prevention Rule in 2016 to reduce methane losses from oil and gas wells on federal
land, but many aspects of that rule have also been repealed recently.
The intention of the 2016 EPA action was to decrease methane emissions by 300,000 short tons (unit from
original citation − amount = 0.27 Tg) in 2020, and by 510,000 short tons in 2025 (0.46 Tg), compared to a
baseline estimate. In contrast, under the revised 2018 EPA proposal, U.S. emissions were perversely expected
to increase by 380,000 tons (0.34Tg) between 2019 and 2025 (Reuters, 2018a): In effect this emits future U.S.
energy security to air. Interestingly, this 2018 EPA proposal was opposed by the industry, which is well
aware of costs: An Exxon Vice‐President stated “We support maintaining the key elements of the underlying
regulation, such as leak detection and repair programs” (Reuters, 2018b).

4.2. Urban Gas Leaks


Urban gas leaks are major sources of methane emissions in most long‐urbanized societies (e.g., Lowry et al.,
2001; McKain et al., 2015; Zazzeri et al., 2017). Many cities in developed nations have built inventories of
estimated methane emissions. However, using aircraft‐mounted observations downwind of major U.S.
cities, Plant et al. (2019) found that in current inventories, natural gas emissions were significantly underes-
timated. Moreover, in many places leak detection is still primarily by human smell. Although methane is
odorless, this is effective as a safety measure because urban gas typically has the organosulfur compound
mercaptan (CH4S) added, which can be detected at a 0.14‐ppb threshold by the human nose (Committee
on Acute Exposure Guideline Levels, 2013).

NISBET ET AL. 22 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

In the Boston, USA, urban region, McKain et al. (2015) reported loss rates
to the atmosphere of around 2.7%, as a proportion of the gas used. In the
U.S. Los Angeles basin, He et al. (2019) found that about 1.4% of commer-
cial and residential gas consumption is released into the air. Such loss
rates are expensive, locally potentially dangerous, and an unnecessary
greenhouse emission. Yet currently many jurisdictions pay little attention
to mitigating urban gas leaks: for example, in the United States, Hopkins
et al. (2016) concluded that “current mitigation approaches are absent or
ineffective.” Even safety can be neglected. A recent example is the series of
explosions and fires in the Merrimack valley, Massachusetts, USA, in
September 2018 that killed one person and damaged over a hundred
buildings (NTSB, 2018). Surveying the urban distribution network in
Boston, USA, Hendrick et al. (2016) found that 15% of leaks surveyed were
potentially explosive, and even small leaks could not be disregarded as
“safely leaking.”
As with gasfields, there is strong evidence that a few major leaks consti-
tute the bulk of emissions from urban distribution networks and that
there is a “fat‐tail” distribution of leaks. In Boston, 7% of leak sources
contributed 50% of emissions (Hendrick et al., 2016) and in U.S. cities, it
is estimated (von Fischer et al., 2017) that repairs to the largest 20% of
pipeline leaks would cut losses by half. Leak prioritization by mobile
measurement is now rapid, inexpensive, and effective (Weller et al., 2018).
Major leaks stand out strongly in vehicle surveys: Their identification does
not need sophisticated quantification. Around the Royal Holloway
location in outer London, gas governor stations, where pressures are con-
Figure 11. Typical source facility: gas processing plant, Hong Kong. Photo:
trolled for domestic distribution, are particular point sources of emissions.
E.G. Nisbet.
It is likely that much could be done to cut emissions by identifying such
specific “likely to leak” parts of the distribution network as mitigation
targets. There are strong grounds (Hausman & Muehlenbachs, 2016) for emphasizing leak reduction over
pipeline replacement as a first option in mitigating emissions from a gas distribution system, especially with
the marked advances in leak detection.
Leak quantification will help optimize mitigation: Identifying and repairing the worst emissions is an attrac-
tive and potentially profitable first step. There is indeed evidence that in the U.S. distribution pipeline leaks
are being reduced (Lamb et al., 2015). However, it is difficult for gas distribution companies to make the
jump between leak detection and leak quantification, and their criteria for repair or replacement may have
different priorities (i.e., safety and proximity to structures, or cost, rather than greenhouse gas emissions).
Currently, it is possible that widespread practice may be to mitigate methane leaks for safety reasons, but
not to be overly concerned about greenhouse implications.
In addition to leak reduction, the replacement of century‐old iron pipes is urgently and widely needed in
Europe and the United States. This is necessary to cut greenhouse emissions and also to improve safety,
by reducing unexpected catastrophic leaks, although the replacement process itself can be hazardous
(NTSB, 2018). Robotic autonomous systems offer much (Robotics, 2017). The use of robotic systems removes
the need to dig up the pipe to seal it. These systems can travel along pipes, detecting leaks and in principle
also sealing many from the inside, including those in “live pipes filled with gas, and potentially giving longer
lifetimes to olds iron pipes.”

In the United Kingdom, replacement is typically achieved without major excavation by inserting new
narrower‐diameter higher‐pressure piping within the older wider‐diameter iron pipes. However, in coun-
tries with regulated local gas monopolies any investment in leak reduction increases the capital base of
the gas industries and thus increases allowable gas prices. Hence proposals from industry to invest in new
pipes may be blocked by regulators. For example, in the past the U.K.'s Office of Gas Supply (Ofgas) resisted
environmental objectives (Danby, 1998). To keep prices down, U.K. regulators restricted corporate proposals
for investment in leak reduction by pipeline renewal. While this may have been politically advantageous in

NISBET ET AL. 23 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

the short term, this extra leakage would have proportionately slightly shortened the useful life of the UK's
domestic North Sea gas supplies, quite apart from the environmental damage.

4.3. “Habitual” Emissions From Domestic and Industrial Facilities


“Habitually” high methane air commonly exists in many places, ranging from near vents of domestic boilers
and water heaters to the surrounds of industrial gas handling facilities such as compressor stations and gas
distribution centers (Figure 11) (Schwietzke, Pétron, et al., 2017; Subramanian et al., 2015; Zavala‐Araiza
et al., 2017). In dealing with the high numbers of small leaks and emissions, low‐cost sensors are becoming
capable of detecting methane anomalies at the sub‐ppm level (e.g., Collier‐Oxandale et al., 2018). Just as
smoke detectors are now ubiquitous, so methane sensors could be widely deployed to identify leaks rapidly.
Domestic heating systems emit methane on ignition because the first gas is only partly combusted. In the
U.K. domestic boilers emit when they ignite for house heating and hot water. In some nations (e.g.,
Australia) “instant” hot water heaters that rely on burning gas are common: they emit methane each time
the hot water tap is turned on. When a gas‐burning system such as a domestic boiler or industrial compressor
switches on or off, the combustion chamber is vented to prevent explosive build‐up of unburnt gas. In
addition, steady‐state burning itself is inherently incomplete. In order to ensure all gas is oxidized and no
stray gas leaks out, down to the ambient background level, the amount of oxygen in the system must be
increased well above the stoichiometric level. The reduction in overall combustion chamber temperature
caused by this excess airflow reduces the boiler efficiency by more than the energy cost of venting unburnt
gas. Because unburned gas passing through can be significant when burners are first turned on, an alterna-
tive strategy to cut emissions may be to pass initial exhaust gas back into the burning chamber for a second
combustion. To mitigate emissions, not all the methane needs to be removed: the challenge is to devise an
inexpensive, robust system to reduce at least some of the excess methane in the exhaust.
Catalytic removal of a significant fraction of excess methane is feasible. Removing harmful trace gases is com-
pulsory for vehicles: In a comparable way, in future methane removal from exhaust gases may become an
attractive and feasible option for stationary gas‐using installations. Catalytic reactors can be autothermal
at methane mole fraction as low as 0.06% but for greenhouse mitigation the catalyst would need heating
(Su & Agnew, 2006). Noble metal catalysts using noble metals such as Pt, Pd, or Rh are effective (Jiang
et al., 2018) and in principle it should be possible to take advantages of economies of scale by using modified
vehicle catalytic converters. But costs may be prohibitive. Gosiewski and Pawlaczyk (2014) estimate that 0.5 t
of Pd would be needed to mitigate emissions from a single coal mine ventilation shaft. Unfortunately,
currently, there has been little research on ultralean methane combustion using these catalysts (Jiang
et al., 2018), though palladium‐based zeolite catalysis is possible (Petrov et al., 2018). More generally, the
use of zeolite‐metal technologies to remove methane from air is promising (Jackson et al., 2019).
Inexpensive metal‐oxide regeneratable catalysts (Buciuman et al., 1999; Döbber et al., 2004) have long been
used in laboratory zero‐air generators in reversible flow reactors. Typically, these reactors use catalysts such
as hopcalite (a mixture of copper and manganese oxides, sometimes with cobalt or other metals), which work
well at moderate temperatures (below 600 °C). These catalysts are typically rapidly made ineffective by water
but can be easily regenerated by heating. Luo et al. (2012) showed full combustion of methane at 575 °C using
biomorphic CuO‐ZrO2 catalysts synthesized on cotton templates. There is also great potential for the direct
catalytic conversion of methane to methanol at low temperature (~220 °C or less) over zeolites that have been
copper‐exchanged (Narsimhan et al., 2016). UV‐photocatalytic removal of methane may also be feasible
(de Richter et al., 2017; Graetzel et al., 1989).
As methane catalysis only occurs at temperatures above several hundred degrees, a vital consideration in
catalytic methane reduction is ensuring that the global warming caused by the extra energy demand needed
to operate the catalyst is smaller than that of the mitigated gas, that is, that there is net benefit. For simple
estimation the heat capacity of 1 m3 of air is about 1.2 kJ. If a catalytic converter raises the exhaust tempera-
ture by 250 °C before venting then the converter must reduce the exhaust methane concentration by well
over 10,000 ppm (assuming heating at 0.1 g CO2eq per kJ). This should be compared to exhaust concentra-
tions from domestic solid‐fuel boilers of 10–100 ppm (EU Ecodesign Directive 2009/125/EC).
One solution to this is to place the catalyst inside the combustion chamber where temperatures are already
sufficient for catalysis and exhaust gas is still hot as it passes over. Vaillant and Gastec (1999) have

NISBET ET AL. 24 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

demonstrated such a system (Figure 12) and achieved 0‐ppm exhaust emissions, an improvement even over
ambient mole fractions. An alternative catalytic approach is to replicate the natural oxidation of atmospheric
CH4 by OH radicals—this is discussed further in section 6.
Finally, it is worth remarking that a more elegant solution to the problem of gas emissions from domestic boi-
lers would be to eliminate the domestic gas boiler and its pipeline supply system entirely. Renewable energy
sources such as wind and sunlight are sporadic. Inevitably, in power systems with high proportions of
renewable generation, there are episodes of superabundant electricity, for example, at 3 am on a windy night,
or on a quiet very sunny afternoon. As decarbonization proceeds, there will be many times of surplus supply
at periods of low demand. Moreover, the costs of nuclear power are typically dominated by capital cost, and
running costs are comparatively small. Thus, nuclear electricity can also be very inexpensive at 3 a.m.
Gas is primarily used either for domestic heating or for electricity generation. In European countries domes-
tic central heating based on circulating hot water is common and wind, and solar power are widely and
increasingly used, coupled with a baseload of nuclear power. Here, rather than using on‐demand gas boilers,
it is likely to become preferable to use surplus‐period (e.g., postmidnight) electricity to heat water, and then
release the heat into the home over the next day and evening. Hot water can easily be stored in an insulated
tank for several days without losing much heat. Cold can obviously be stored also: In hot places it may
become preferable to operate air conditioning by blowing air over a thermal mass such as ice chilled at a time
when electricity was abundant. The intrinsically leaky urban gas pipeline network could be repurposed as
cabling conduits both for telecommunication and, as vehicles turn electric, for electric power supply.
Closing the domestic gas network and switching to renewable‐generated electric heating may have
considerable synergy with the move to battery‐powered electric cars. In the United Kingdom it would benefit
long‐term energy security if gas imports were halted and the declining local North Sea gas reserves were
retained solely for use in peak‐demand electricity generation, rather than used up for domestic heating.
4.4. Urban Wastewater and Sewage
Sewage systems host anaerobic methanogens and thus emit methane at all stages, from house drains to was-
tewater treatment plants (Liu, Ni, et al., 2015; Guisasola et al., 2008, 2009; Fries et al., 2018). Where biogas is
extracted, it may be incompletely combusted. Drains are prone to methane explosions, while methane pro-
duction in the sewage reduces the readily biodegradable chemical oxygen demand (Guisasola et al., 2008),
detrimental to processes in the wastewater treatment plants. Thus, methane emissions from sewage consti-
tute a safety and processing problem, as well as being a greenhouse concern. The wider problem of methane
removal from sewers in the long gravity runs from buildings to wastewater treatment plants is potentially
addressable by installing small bioreactors at intervals along the length of the sewers, and by extracting
methane‐rich air above the sewage for catalytic or soil methane oxidation.
Various techniques can be used to reduce emissions, but at present efforts have focused on the wastewater
treatment plants. Here, methane recovery efficiencies over 50% can be achieved, for example, via a
submerged underwater bioreactor (Giménez et al., 2012). Floating bioreactors on wastewater settling pools
have achieved 67% methane oxidation in agricultural settings (Syed et al. (2017), and the technology may be
transferrable to urban wastewater. Matsura et al. (2015) achieved over 99% removal of dissolved methane by
a two‐stage system using a downflow hanging sponge reactor to remove methane from an upflow of anae-
robic sludge effluents. Similarly, using an upflow anaerobic sludge blanket reactor at ambient temperatures,
Bandara et al. (2012) achieved good removal efficiencies of more than 50% in summer, as measured by
chemical oxygen demand, though less than 40% in winter. No biogas was evolved. More generally, methane
removal in wastewater treatment plants appears to be both feasible and inexpensive, partly self‐financing in
both reduction of safety risk and perhaps in providing methane for ancillary power.
4.5. Methane Emissions From Biodigesters and Biogas Generators
Biogas and oils are increasingly being produced by anaerobic digestion of various substrates such as domes-
tic and industrial food waste, manure, or sewage sludge. There are significant benefits from this activity,
including energy generation that replaces fossil fuel combustion, and mitigation of emission of methane
from manure. However, as biogas is typically 50–70% CH4 and 30–50% CO2, if part of the biogas is lost to
the air, then there is a risk of cutting CO2 emissions but increasing net greenhouse warming impact by
the emission of methane.

NISBET ET AL. 25 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 12. Concept for completely catalytic boiler system. The premixed gas air mixture is distributed over a heat exchanger. Modified on a concept from Vaillant
and Gastec (1999).

The production of renewable energy from biogas plants is growing rapidly worldwide. In the United
Kingdom, biogas plant numbers have increased almost 500% in the past 5 yr, with 607 operational biogas
plants in 2018, and around 500 new plants being considered for construction (ADBA, 2018). The European
Biogas Association anticipates that overall biogas production will be at least 50 × 109 m3 by 2030, capable
of providing 2–4% of the EU's electricity needs by displacing methane from fossil fuels (EBA, 2019).
Unfortunately, biogas plants can have very high fugitive methane emissions. The emission rate depends on
engine construction, plant design, and operation. Emissions can come from water handling facilities, gas
engine exhausts, gas flaring, leaks from pipes, biogas upgrading units, tanks, etc., or from deliberate venting
(Angelidaki et al., 2018; Duren et al., 2019; Fredenslund et al., 2018; Liebetrau et al., 2018; Samuelsson et al.,
2018; Scheutz & Fredenslund, 2019). Some emissions can be single large leaks or long‐lasting bursts from
pressure relief valves (Reinelt & Liebetrau, 2019). In particular, Kvist and Aryal (2019) found that water
scrubbers and pressure relief valves were especially significant sources of emissions, a finding similar to find-
ings around many unconventional natural gaswells. To date biodigester emissions have had comparatively
little attention compared to leaks from other types of gas production facilities, even though biodigester leak-
age rates are likely at time in excess of leakage from fracked gaswells. Typical methane emission rates range
between 0.4% and 15.0% of the total methane production, and overall the methane leaks averaged 4.6% of
methane production (Scheutz & Fredenslund, 2019). These loss rates are broadly comparable to the leakier
end of the fracking range. Reducing methane emissions from biogas plants will probably take the type of
painstaking attention to detail seen on better managed natural gas production facilities, for example, by
reducing flaring, and restricting deliberate venting, and changing the way pressure relief valves operate.
Kvist and Aryal (2019) showed that regenerative thermal oxidation had a major impact, in their study redu-
cing the methane mixing ratio of from 6,900 ppm to an average of 13 ppm and suggested that in jurisdictions
where methane emissions incurred financial penalties, this method of removal would become attractively
advantageous to plant operators. Much can be done.

4.6. Landfill Emissions, Particularly in Less‐Developed Countries


In addition to natural gas systems, large urban methane emissions also come from urban landfills. Saunois
et al. (2016) cite bottom‐up evidence for global emissions around 60 Tg annually. Landfill emissions can be
readily mitigated, either by capture and piping, or simply by adding a thicker soil cover to act as host habitat
for consortia of methanotrophic bacteria.
In some developed nations, particularly in Europe, landfill emissions are now heavily regulated. Figure 2
shows the effectiveness of this policy measure. Since 1990, U.K. annual emissions of methane from landfills

NISBET ET AL. 26 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

have been reduced to about a quarter of the previous amount (UK NIR, 2019). Worldwide, however, landfill
regulation is weak, even in some jurisdictions with strong capabilities, such as parts of the United States.
For example, Duren et al. (2019) found that landfills are the largest emitters of methane in California and
dominate methane point‐source emissions in the state (41% of total). Although such emissions are not
well reported in state and national inventories (Duren et al., 2019), they are clearly targets for mitigation.
Ideally, waste inputs should be controlled, with organic wastes diverted to composters and biodigesters, and
landfills themselves should be well lined and covered with geomembrane, and typically fitted with gas
capture piping, delivering gas to fuel electricity generators. If the methane itself cannot be extracted and
combusted to make power, biofiltration is effective in removal (Gebert & Gröngröft, 2006; Park et al.,
2009). But such systems require both investment and ongoing commitment to management as pipes can
frequently fracture and lead to large unidentified localized leaks. Moreover, linked facilities such as
anaerobic biogesters can themselves be very high emitters of methane (see section 4.5). In particular, the
working face of the landfill, which is where emissions are most active (Figure 13), necessarily remains
uncovered, with the gas uncaptured.
Nevertheless, active mitigation of landfill emissions in developed nations is a long‐established and effective
priority and thus not the focus of the discussion here. In contrast, in many tropical nations, landfill methane
emissions are wholly uncontrolled and are characteristically high, especially in India (e.g., Ghosh et al.,
2019; Kumar et al., 2004; Lobert et al., 1990). Even in Kuwait, the landfill is a major source (al‐Shalaan
2019) (Figure 3). There is often little interest in investment in methane emission control and there is poor
management of the emissions in many cases. Fires often break out, either accidentally or deliberately. For
example, South Africa, although it has strong landfill regulation on paper, has widespread illicit burning
(Roberts, 2018). In less regulated tropical and subtropical locations waste is widely burned (Figure 14) and
in many cities daily large scale burning of local waste piles equivalent to landfill occurs and is likely to emit
both significant amounts of methane and much harmful air pollution.
Landfill soil covers host active consortia of methanotrophic bacteria, and thus, bacterial methane oxidation
is an attractive option for low cost methane mitigation (Scheutz et al., 2009). In U.S. landfills with 0.3‐ to
1.2‐m thick soil covers, typically clay, Chanton et al. (2011) showed that a third or more of the egressing
methane (37.5 ± 3.5%) is oxidized by methanotrophs in these covers. While the uppermost soil crust may
be dry, with depth the oxygen and nitrogen mole fraction decline and methane and CO2 increase, especially
in active faces and in landfills that are not membrane covered. Stable isotopes can be used to assess the
impact of methanotrophy (Sparrow et al., 2019). Methane emitted from below passes upward into increas-
ingly aerated soils, and methanotrophic oxidation rates in moist but not waterlogged soil can be extremely
high (Whalen et al., 1990). Methanotrophic oxidation has been suggested by Stein and Hettiaratchi (2001)
for mitigation around heavy oil wells and also sanitary landfills in Alberta, Canada. Moreover, soil oxidation
can be effectively enhanced by simple techniques, such as enrichment with biochar (Reddy et al., 2014;
Yargicoglu & Reddy, 2017).
The usefulness of such passive biocover systems was demonstrated by Scheutz et al. (2014), who studied an
old unlined landfill, comparable to modern tropical landfills. They used a bioactive cover made of local
compost materials such as garden and kitchen waste, and demonstrated an average mitigation efficiency
of around 80%. Although the landfill they studied is in Denmark, even in freezing winter conditions the
exothermic bioreactions maintained a high mitigation efficiency in the system, and optimal biosystem
temperatures of 30–40 °C. This is because, although methane oxidation is strongly temperature dependent
(Wang et al., 2011), landfills are strongly exothermic. Einola et al. (2007) found that exothermic methane
consumption continues in landfill cover soil even in cold meteorological conditions, a finding that may have
value for mitigation in cool‐winter regions such as north China, central Asia, and high‐altitude Bolivia
and Peru.
Significant reduction in emissions, up to 96% in summer (Lee, Jung, et al., 2018), can be potentially achieved
by inexpensive soil or compost coverings. Financially, Scheutz et al. (2014) showed that simple biocover
mitigation was cost‐effective: Indeed, it is possible that in some locations such as the Mediterranean it
may be viable to use carbon credits to finance soil‐cover landfill mitigation directly (Kormi et al., 2018).
Moreover, soil management of emissions also contains unwanted smell and fire risks and is likely to gain
local support (especially if the matured cover vegetation can then be given over to urban parks).

NISBET ET AL. 27 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 13. Vehicle‐mounted mobile survey of U.K. landfill, superimposed on prefill topography. Note very high ambient methane downwind of active cells. Map
data: Google DigitalGlobe.

Although methane mitigation in landfills in developed nations is typically complex and demands strong
investment and monitoring, for less wealthy urban regions there is a wide array of simpler cheaper options
to mitigate emissions that do not demand high capital investment or complex skills, and which can poten-
tially be assisted within development aid budgets. Thus, where landfill gas extraction by complex piping is
prohibitively capital intensive, soil biocovers teamed with careful attention to rapid covering of recently
active faces offer high prospects of containing and mitigating emissions. Mitigation of emissions from com-
posted heaps and from residual landfill can be achieved by methane reduction via bacterial oxidation in
the soil covering the landfills (Abushammala et al., 2014; Boeckx et al., 1996; Sadasivam & Reddy, 2014;
Serrano‐Silva et al., 2014). In addition, in nations with high unemployment it is not expensive to adopt waste
sorting for compostable materials.
With the rapid growth in tropical urbanization, methane emissions from tropical cities are probably now
very large. Although they are very poorly studied and quantified, tropical urban emissions may be becoming
significant in the global budget, and thus a mitigation target of growing importance. In particular, where

NISBET ET AL. 28 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 14. Waste heap fire in Pomona dump, Harare, Zimbabwe, November 2016. (Photo: T. and L. Broderick, Project
MOYA).

regulatory capacity is limited, rather than attempt expensive and complex landfill gas recovery, as an initial
mitigation method it may be more effective in newly urbanizing countries to adopt simpler strategies: cover
the landfills with soil and sort the waste.

4.7. Deep Coal


Methane emissions from underground coal mining activities are important contributors to the global
methane budget. They occur both when vented from coal mining and associated surface coal processing
and from leaks from coalbed methane extraction (Moore, 2012). China, with the world's largest coal indus-
try, is especially important (Thompson et al., 2015), with methane emissions dominantly from underground
operations (Zhang & Chen, 2010). Ju et al. (2016) estimated China's underground coal mine emissions in the
year 2007 at about 19 Tg.
Figure 15 shows mapped methane mole fractions around two U.K. coal mines, with well over 5‐ppm
methane widely present in places, especially coal piles and mine vents. Methane comes both from thermo-
genic (isotopically heavy) methane associated with the coal, and young microbial methane by reduction pro-
cesses during and after mining (Krüger et al., 2008). Air ventilated from coal mines generally contains
methane, often with mole fractions so high it is explosive or can be burned. Such methane‐rich air, which
can be removed by combustion in thermal reactors (Karacan, 2011), has always been of immediate safety
concern to all coal mines.
Underground mine vents are emission hot spots, as shown in Figure 16, making such emissions a promising
target for mitigation. Emissions estimates from Australian coalfields demonstrate that targeting the reduc-
tion of methane emissions from underground coal mine venting could yield significant mitigation benefits.
In the State of New South Wales, Australia, there are 19 underground coal mines that produce 24% of the
annual mined coal in the state (https://www.coalservices.com.au/mining/statistics). However, these vented
emissions from underground coal mines account for 82.6% of the coal sector's reported emissions at the state
level, 30.7% of the State's total methane emissions, and 8.4% of Australian national methane emissions
(Australia's National Greenhouse Gas Accounts, 2017; http://ageis.climatechange.gov.au). They are an
obvious target for mitigation.
Mitigating methane emissions from deep coal mines is not necessarily expensive. Holmes (2016) describes a
12‐month trial involving comprehensive emission reduction in the Australian Hunter Valley coalfield, based
on vigorous efforts to locate and seal leaks and to use measures such as pressure balancing to suppress egress.
This work, entirely practical and using common sense, likely brought marked safety improvements and had
little capital cost; it removed about 4 kt of methane or 80 kt of CO2 equivalent (quantities and units cited as in

NISBET ET AL. 29 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 15. Google Earth view of methane mole fraction columns (red contours) measured along the transect downwind of Aberpergwm and Unity deep mines in
Wales on 17 October 2013. The yellow line represents the 5 ppm mole fraction level, yellow markers samples location. Map data: Google, DigitalGlobe.

Figure 16. Mobile mole fraction methane measurements in the midcoast and Hunter coalfields of eastern Australia, recorded in 2014 (University of New South
Wales and Royal Holloway, University of London). The background methane mole fraction for the region was 1.78 ppm. Emissions from the underground coal
mines exceeded the measurement range of the analyzer (20 ppm), while throughout the Hunter coalfield the average methane mole fraction was 2.02 ppm. Map
data: Google, DigitalGlobe.

NISBET ET AL. 30 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

the original source, Holmes, 2016, which used a Global Warming Potential of 21), for a cost of Aus$1.28
($0.9USD) per ton of CO2 equivalent. Such efforts are simple good management and should easily be imple-
mented in a safety‐conscious setting or with effective regulatory oversight.
Both working and closed coal mines also emit methane‐rich air, well above ambient mole fractions but well
below explosive‐risk thresholds, which may not be of concern to mine managements on safety grounds but
are important in terms of greenhouse gas emissions. This air has methane content that is too low to sustain a
thermal reactor, but as with boilers and compressors, emissions can be mitigated in catalytic reactors.
However, usually, the air is moist and the water degrades the catalytic effectiveness: Thus, the flow needs
to be reversible to regenerate the catalyst regularly. For nonexplosive mixes, reverse flow reactors are effec-
tive in mitigating methane emissions from coal mines for air with ~0.3% methane and 5% water (Fernández
et al., 2016). However, for mitigation purposes, air with methane as low as 5–10 ppm would be a target.

4.8. Open‐Cast Coal


Open‐cast (or open‐pit or open‐cut) coal mines are also a major source of emissions (Dontala et al., 2015). In
contrast to deep coal, where the methane egresses from localized vents and openings, open cast mines have
spatially spread‐out emissions, which may be greatest near the active excavation face. Moreover, these emis-
sions can be isotopically light, comparatively rich in 12C, indicating such methane is not thermogenic but is
young, biologically made by in situ methanogens, in anaerobic settings (Zazzeri et al., 2016).
As with landfills, managing emissions from open cast coal mines should be possible. Many open cast mines
currently remove the cover from many kilometer lengths of coal beds and then strip it in kilometrs long
benches, exposing large areas of uncovered organic‐rich benches to develop anaerobic conditions just below
the remaining surface. In contrast, landfill management seeks to expose the active face only, and rapidly to
cover all inactive cells with aerobic soil that hosts methanotrophs. Changing open cast coal pits to short
active benches and covering inactive surfaces quickly with methanotroph‐hosting aerobic soil cover would
likely reduce emissions substantially. More complex mitigations could be to ensure that the coal face is well
aerated, and thus aerobic. It is also intuitively possible that arrays of high‐powered UV‐LEDs could be used
to kill any surface methanogenic bacteria, but this has not been tested. Such changes would demand corpo-
rate cultural and management innovation and redesign, but not necessarily any major capital costs although
the energy cost of the implementation must be taken in account to demonstrate net GWP reduction.
In the medium term, it is perhaps more likely that the entire coal industry will become uneconomic, except
for coal destined for steel production.

4.9. Methane Created by Anthropogenic Biomass Burning in the Tropics


Emission of methane from deliberate biomass burning is globally important, mostly from tropical C4
savanna grasslands and crop waste and also from boreal and tropical forests. In managed boreal and dry tro-
pical forests, fire suppression is controversial and the accumulation of leaf‐fall and branch‐fall fuel load may
simply eventually generate larger fires. In the tropical C4 grasslands, however, under the seasonal shifts of
the Intertropical Convergence Zone, fire suppression can be very advantageous, preventing valuable soil
nutrients being lofted in the smoke plume and eventually deposited far away in the oceans.
Fire emissions are especially important in Africa (Andela et al., 2019; Andela & van der Werf, 2014; van der
Werf et al., 2017; Zubkova et al., 2019), where burning occurs very widely in savanna grasslands (Figure 17),
and in leaf litter fires in facultatively deciduous dry savanna woodlands. Burning can be accidental when
crop waste fires spread into neighboring untended communal grazing land, or deliberate to bring on new
green shoots of regrowing perennial grasses. Burning is also important in African cropland to clear crop resi-
dues, and as waste plastic burning.
As an example, at Bachok Marine Research Station (BMRS) on the NE coast of peninsular Malaysia, compel-
ling evidence of burning emissions correlates with periods when the sea‐breeze breaks down and land
breezes dominate (Figure 17). During such episodes the methane mixing ratios at the station rapidly increase
over a matter of minutes, along with increases in anthropogenic tracers such as CO2, volatile organic com-
pounds, and NO2 (Dunmore et al., 2016), as well as combustion markers such as carbon monoxide (CO) and
acetonitrile (CH3CN) (e.g., Lobert et al., 1990). Although methane emission in Malaysia is more complex
than this simplification (with isotopes of methane during land breezes suggesting a mixed source

NISBET ET AL. 31 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 17. Continuous methane (CH4) and acetonitrile (CH3CN) measured over 3 days at Bachok station, Malaysia. The
inlet height is 30 m, and the site is located 50 m inshore. The correlation between enhancements of CH4 and CH3CN is
indicative of a strong burning contribution to the CH4 enhancement (unpublished results, University of East Anglia/
University of Malaya).

including a biogenic source), it is clear that methane emission from all forms of waste disposal (rural
agricultural crop waste, village debris, and urban waste) is widely underregulated, with unquantified
amounts of methane entering the atmosphere. Such fires, in addition to producing methane, CO,
particulates and black carbon, deplete essential soil nutrients such as nitrate and impoverish soil organic
carbon, exporting them to the smoke.
Fire numbers vary strongly with meteorology and can peak during and after major El Niño events (Chen
et al., 2017). Fuel loads are also strongly dependent on preceding meteorology (Zubkova et al., 2019).
Satellite data sets may show a recent decline in global burnt area (Andela et al., 2017), and globally,
Worden et al. (2017) showed evidence of a decline in biomass burning between the 2001–2007 and 2008–
2014 time periods, which may reflect complex socio‐economic changes as much as meteorology, though
Forkel et al. (2019) found the trend is not significant. Zubkova et al. (2019) concluded that changing climate
patterns and increased terrestrial moisture facilitated a decline in African fires in the 2002–2016 but pointed
out that most African fires are human‐caused and the fire‐climate‐human relationship is complex.
Wood and dung fuel burning is also widespread in the tropics, for cooking and heat, but slowly being
replaced by electrification, which is progressing across most tropical nations. In India, widespread crop resi-
due fires have been blamed for air pollution and may soon be controlled. In the savanna regions of Australia
fire management is used to obtain carbon‐offset units that may be sold to generate additional income, redu-
cing the overall carbon emissions by doing carefully managed cool burns at the beginning of the dry season,
rather than letting uncontrolled hot burn happen at the end of the dry season. Cool burns release less carbon
dioxide, and the early dry season burns reduce the fuel available for uncontrolled hot burns that can natu-
rally occur at the end of the dry season. This may reduce CO2 emissions but not necessarily CH4, which is
emitted under partial combustion conditions and may be increased with a moister fuel load (http://www.
cleanenergyregulator.gov.au/ERF/Choosing‐a‐project‐type/Opportunities‐for‐the‐land‐sector/Savanna‐
burning‐methods).

NISBET ET AL. 32 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Land management is crucial in controlling burning. In Zimbabwe, for


example (West, 1971), the commercial farmland, partly wooded and
partly C4 savanna grasslands, was divided for many decades into
“Intensive Conservation Areas” where vigorous fire suppression was prac-
ticed. As a result, soil nutrients were conserved and soil carbon built up.
Savanna Africa has had tens if not hundreds of millennia of grass burning,
both for crops in recent millennia and prior to that for hunting.
Consequently, the region has developed a very widespread anthropogenic
fire ecology (West, 1971). Suppressing burning clears the air and reduces
methane emissions, but it changes the vegetation and the animals: fewer
lion, zebra, and buffalo, more bush pigs, monkeys, and rats (West, 1971).
With the abandonment of most commercial farming in the early 2000s,
fire suppression halted in Zimbabwe. Interestingly, however, in recent
years, regrowth of fire resistant trees has taken place in the unfarmed
areas (own observations), and thus, C4 grass fires (Figure 18) may be
eventually replaced by less intense C3 leaf litter fires.
Figure 18. Grass fire, August 2016, Harare, Zimbabwe. Photo: Lucy
Broderick. Agricultural practice is widely changing, and increasing awareness of fire
nutrient losses is likely. Widespread air pollution from smoke plumes that
can be tens to hundreds of kilometers long can present major health hazards. Fire suppression is feasible and
widely advantageous. Personal observation and discussion in Zimbabwe's communally held lands (EGN)
suggests that annually burnt areas may be substantially reduced by small‐scale local reward programs that
could reward schools or services in communities that had not burned their communal grazing land. Such
programs could be validated inexpensively by end‐of‐season satellite observation. Public policy has also dri-
ven fires: For example, the European Union's demand for palm oil biodiesel has played a role in burning for-
est in SE Asia to convert the land to palm oil plantations (Obidzinski et al., 2012; Susanti & Maryudi, 2016). Is
it wise to reduce Europe's CO2 emissions by deforesting Indonesia? Thus, there is room for optimism that
tropical biomass burning may be sharply reduced in the near future, relatively inexpensively, provided major
climate events such as drought or heat waves do not intervene.

5. Emission Removal—Intractable Emissions


In many cases, emissions are regarded as “intractable” (i.e., cannot easily be stopped). Yet these need to be
addressed. These emissions are typically disseminated, from many sources, some moving (e.g., animals).
Such emissions are often regarded to be “inevitable” impacts of essential human activities like food produc-
tion and occur where the mole fractions in ambient air are too low to consider combustion. Funding agen-
cies have neglected the problem of methane extraction from ambient air, and where methane mitigation has
been targeted, the focus has understandably been on emission reduction. In particular, the possibilities of
methane removal on farms have had little attention. Currently, many agricultural emissions of methane
can be seen as too intractable to consider for affordable mitigation.
In principle, as in landfills, methane removal around intractable sources can be achieved via soil methano-
trophy. The movement of methane into the soil is controlled by gas diffusivity, which depends on the soil's
air‐filled porosity, and the capacity of the soil for uptake of methane depends on soil moisture and aeration
(Ball et al., 1997; Smith et al., 2018). Aerated soils such as sandy loams host methanotrophs, capable of living
on the ~2 ppm methane mole fraction in ambient air, and thriving when the mole fraction is higher.
Coupling biofiltration (Leson & Winer, 1991) with enhanced methanotrophy (Jiang et al., 2010; Yoon
et al., 2009) could provide a powerful approach to a wide range of problems.
Afforestation has been much discussed as an attractive option for CO2 capture (e.g., Humpenöder et al.,
2014), but there has been less attention to the usefulness of afforestation in methane removal. Yet methano-
trophy in forest soils can be very effective in taking up methane. Wu et al. (2018) showed that in central
China the reforestation of cropland could create an important methane sink. They showed that recreating
woodland nearly doubled methane uptake, and that shrubland, which took up ~37 μg·m−2·hr−1 (around
3.2 kg·ha−1·yr−1), had nearly triple the methane removal of cropland. In natural settings, methane oxidation
is highest in well‐drained upland soils (Christiansen et al., 2016). More generally, coarse‐textured moist but

NISBET ET AL. 33 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

not waterlogged forest soils show the most rapid methane oxidation rates (Levy et al., 2012; MacDonald
et al., 1996), especially when the air can easily diffuse through the surface litter and the soil has a well‐
developed structure (Smith et al., 2003). Wide areas of the tropics, for example, in Africa, have been
deforested in the past century and in many areas replaced by poorly yielding cropland. The primary reasons
for reforestation are land restoration, improvement in perennial streams and carbon sequestration, to which
the possibility of methane removal provides a small but useful added incentive.

6. Reducing Agricultural Methane Production


Agricultural methane emissions are very large, globally over 130 Tg/yr, with over 100 Tg/a from enteric
fermentation and manure, around 30 Tg/a from rice cultivation, and in addition about 30 Tg/a from biomass
burning, mostly deliberate (estimates from Saunois et al., 2016). This review journal focusses on specifically
geophysical topics and the task of reducing agricultural methane emissions is a topic too large to be
discussed here in detail: There is space only for a brief synopsis. Thus, the topic of methane reduction by
changing pasture management and animal and human diets is not covered in detail here.
There is a very large literature on this (e.g., Benchaar et al., 2001; Buddle et al., 2011; Smith et al., 2008) but
that is outside the scope of this review. The problem is challenging. Methane emissions from rural agricul-
tural sources such as rice fields, cattle, and biomass burning (much of it deliberate to clear crop wastes or
fields before planting) tends to be from widely disseminated or geographically spread out sources, sometimes
in different locations from day to day or even hour to hour.
Integrating climate change mitigation goals into agricultural practice is urgent but poses great difficulties
(Campbell et al., 2017; Fellmann et al., 2018). Cattle in the U.S. and Europe emit up to 400 g/day of methane
(for an assessment see Niu et al., 2018). In particular, cattle feedlots and barns can be superemitters, as
reported here using University of New South Wales/Royal Holloway, University of London measurements
around Beef City, a 26,500 cattle feedlot in Australia (Figure 19). Similarly, Duren et al. (2019) found that
dairies alone accounted for 26% of California's methane point‐source emissions. Given the size of the total
emissions, even small percentage reductions will have large impacts.
Methanogenic archaea are strictly anaerobic but find good habitats in the digestive systems of oxygen‐
breathing land ruminants. Methane is breathed out from all foregut fermenters—cattle, sheep, goats, and
deer. Note that cattle primarily produce methane from the front end—bovine eructation dominates over
flatulence. Moreover, it is not only the breath of animals that produce methane: If feces become anaerobic,
then archaeal methanogenesis takes over. Thus, cattle pats in dry tropical grassland pastures are unlikely to
produce much methane, nor do piles of elephant dung (RHUL observations), but large manure tanks in
industrial farming are major sources (Veltman et al., 2018).
Some nonruminants are also methane emitters, although not on the scale of cattle. Hippos, camels, and
llamas also process cellulose in similar ways, though not strictly ruminants. Hoatzin birds also ferment in
the foregut, as presumably did dinosaurs in related clades (Wilkinson et al., 2012). Hind‐gut fermenters, such
as horses, elephants, and rhinos, appear to produce much less methane. Humans are a minor source of
methane. Roughly 15–20% of humans breathe out methane (own observations; see also Polag & Keppler,
2019). This methane is presumably created in anaerobic parts of the digestive system and then transferred
from blood to lungs.
6.1. Practical Mitigation of Methane Emissions From Farm Animals
Methane emissions from hard‐to‐mitigate sources, such as enteric fermentation, are very large (Rogelj et al.,
2015; Reisinger & Clark, 2018), too important to ignore as intractable. For example, high methane mole
fractions, often 10‐100 ppm can occur in air around feed lots (Figure 19), cattle pens (Grainger et al.,
2007), manure heaps, agricultural biodigesters (Flesch et al., 2011), and near active faces of landfills
(Riddick et al., 2017; Zazzeri et al., 2015). Methane yield from manure is highly variable, depending on
the diet fed to the cattle (Amon et al., 2007).
In U.S. cases, substantial reductions in methane emissions can be achieved by more sophisticated farm man-
agement (Dairy Coordinated Agricultural Project, 2019). Implementation of a beneficial whole farm man-
agement packages reduced carbon footprints substantially, by between roughly a third to more than 40%.
Manure management is particularly effective, using improvements such as an anaerobic digester with

NISBET ET AL. 34 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 19. Beef City feedlot and processing facility downwind methane plume, Queensland Australia, mapped in 2014
(UNSW and RHUL). This facility processes more than 1,000 animals per day, while the feedlot can hold up to 26,500
head. Map data Google, DigitalGlobe.

methane capture for biogas, and sealed manure storage with gas flares (Veltman et al., 2018). This means
manure tanks, lagoons, heaps, and stores are attractive first targets in efforts to mitigate agricultural
emissions. Widely present, at sites like feedlots, barns and milking stations, they are localized and
contained sources of high‐methane air where emissions can be mitigated by careful manure management
in oxidizing settings, including drying manure, using anaerobic digestion, or by catalytic oxidation (Pratt
& Tate, 2018).
Van der Zaag et al. (2018) investigated two options for reducing emissions from manure—(1) separating
solids and liquids and (2) anaerobic digestion. For solid‐liquid separation they found an 81% average
methane emission reduction compared to raw manure, on a per liter basis. For anerobic digestion, the
reduction was 59%, on average. More generally, in contrast to industrial‐scale diaries with large numbers
of animals in cattle courts, barns or feed lots, preferring “organic” grass‐fed pasture‐based dairy farming,
with rapid oxidation of manure in the field, should also reduce methane emissions.
Globally, most cows live outside the United States and Europe, with the heaviest cattle populations in the
tropics and subtropics, especially in India with 150 million small dairy farmers, and in small‐scale beef
and dairy farming in tropical Africa (Gilbert et al., 2018) where few technical resources may be available.
Here management practices are very different and perhaps less amenable to emission reduction, but
per‐cow emissions may be lower than in industrial farming as cattle are largely free to roam, grass and
browse‐fed, and manure is typically left on grazing lands or collected and burned as fuel.
Although changes in cattle diet and management, such as open grazing on nitrogen‐managed pastures
(Warner et al., 2015) rather than stall feeding, can be effective in cutting emissions, these approaches (not
discussed further here, but accessible in veterinary journals) are often costly, only partly effective, or demand
high skill levels by the farmer. They may be difficult to apply in nations such as South Sudan, with very high
cattle population densities but low governance capacity. Other innovative efforts to use microbiological
intervention, including diet changes, to reduce emissions from cattle and sheep are also not discussed here
as this very large topic is outside the scope of this essentially geophysical review (but see Smith et al., 2008).
Where animals are housed, particularly in winter, air is often extracted, and there is thus an opportunity for
mitigation of methane emissions. Air extracted from intensive pig and poultry production can also generate

NISBET ET AL. 35 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 20. Conceptual model of low‐cost reduction of emissions from cattle. Extractor fan removes air under mouths of cattle in milking shed or winter cattle court.
Air is pumped to biofilter growth medium in a greenhouse, polytunnel, or under plastic field covers. Power from rooftop solar or wind.

high methane mole fractions (Van der Heyden et al., 2015). In Europe some facilities housing cattle also have
air extraction systems installed. Note that methane removal would not have to be complete nor continuous,
merely substantial. The ethical balances in industrial animal farming are not addressed here.
6.2. Practical Mitigation of Methane Emissions From Rice Paddies
Methanogenic habitats are created in wetland rice paddies. Zhang et al. (2016) found the magnitude of emis-
sions depended mainly on the water management procedures. Studying the 2000s decade, they estimated
annual global rice field methane emissions to be under 20 Tg under intermittent irrigation, but nearly 40
Tg under continuous flooding in the 2000s. It is in concept possible to reduce methane emissions from wet-
land rice production, both by directed water management and by selection of rice varietals and application of
fertilizers and organic residues (e.g., Smith et al., 2008). However, as with diet‐management for cattle, wide-
spread reduction of methane emissions by deliberate management of rice production would likely demand
high skill levels and strong governance, and perhaps also acceptance of lower yields. Moreover, in some
cases there can be a trade‐off between methane emission and N2O emission (e.g., Naqvi et al., 2018; Smith
et al., 2008). Net benefit to the greenhouse is not necessarily obtained by cutting methane output. Another
interesting option is the use of biofertilizers to enhance methanotrophy in fields (Singh & Strong, 2016).
6.3. Conceptual Models—Future Possibilities
One approach for dealing with methane emissions from intensive animal farming may be abiotic removal
via ambient‐temperature UV‐photocatalysis of methane and water using TiO2 substrates, in a process ana-
logous to the natural oxidation of atmospheric CH4 by OH radicals (de Richter et al., 2017; Graetzel et al.,
1989). For example, Costa et al. (2012) and Guarino et al. (2008) painted the walls of pig barns with TiO2
paint and hung UV lights. They both measured significantly reduced methane in the barn exhaust.
However, Maurer & Koziel (2019) and Liu, Maghirang, et al. (2015) ran exhaust air through external
catalysis reactors, and neither of these saw a methane reduction. Thus, it is possible the pig barn experiments
did not actually react any methane, but rather they may have UV irradiated and killed the methanogens
in the manure on the floor. A further problem is that OH radicals will oxidize the most reactive molecules
first, and thus methane last (Hay et al., 2015). This means that photocatalysis of methane in barn air
laden with organic molecules is more likely to attack other species before methane: this is good for smell
reduction but not methane removal. Photocatalysis in clean ambient air however is potentially more feasible
(de Richter et al., 2017).
Biological options may in future be more attractive than abiotic approaches for methane removal. Most
biofilter studies have investigated removal of methane at mole fractions in excess of 1,000 ppm (Melse &
van der Werf, 2005; Ramirez et al., 2012), but Girard et al. (2011) report effective use of biofiltration in
removing up to 45% (and potentially up to 65%) of methane emitted from swine slurry in Canada, at mole
fractions as low as 250 ppm (Ramirez et al., 2012). However, significant extra study will be required due

NISBET ET AL. 36 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Figure 21. High methane in ambient air around Zimbabwean cow. Current RHUL study. Photo: Lucy Broderick.

to the extremely different conditions under which a low methane concentration biofilter will operate, and
there is a risk the biofilters may convert ammonia into nitrous oxide, another powerful greenhouse gas
(Melse & Hol, 2017). Yoon et al. (2009) modeled the behavior of trickling biofilters using standard
methanotrophic bacteria and found that operation at ~500 ppm was possible. Potentially, extraction
should be feasible even down to ambient air mole fractions if soil microbes that can consume ambient
concentrations are used; however, rates may be slow and some cells probably cannot survive in a
standard trickling biofilter as the falling water shears cell walls and cell loss through shearing is greater
than cell growth (Yoon et al., 2009). Passive airflow biofilters may offer a more favorable environment.
Unfortunately, despite their apparent promise, one significant concern with using biofilters at low methane
concentrations is that their operation may inadvertently increase other greenhouse gas emissions, in parti-
cular emissions of N2O (which has a lifetime of over a century and a global warming potential 10 times
greater than methane over 100 yr). Specifically, a review of bio‐filters in the swine industry has shown that
biofilters increase the amount of N2O present of up to 400% (Van der Heyden et al., 2015, and references
therein). The proposed biochemical pathway is the conversion of NH3‐N with 10–40% efficiency, with long
residence times of the air in the filter, necessary for low concentration methane removal, positively corre-
lated with emission. It has been suggested that governments should be reluctant to permit biofilters to be

Figure 22. Simple box model to show the potential impact of mitigation. Left‐hand panel shows modeled emissions. Right‐hand panel shows the results in terms of
the methane mixing ratio. Purple line approximates Pathway RCP2.6 (Collins et al., 2013; Rogelj et al., 2012), which is compliant with the Paris Agreement.
Red line assumes starting global emissions of 520 Tg/a, increasing with a step change of 25 Tg/a in 2007–2020, to match the recent observational record (Nisbet
et al., 2019). From 2020, the paths diverge. Blue line—no change in emissions after 2020. Orange line—10% cut in emissions spread linearly over the period
2020–2055, followed by stable emissions. Green line—20% cut over period to 2055, followed by stable emissions. Red line—30% cut in methane emissions to
380 Tg/a in 2055, followed by stable emissions, which cuts the atmospheric mixing ratio of methane to about 1,200 ppb by the Year 2100.

NISBET ET AL. 37 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

installed at livestock operations (Melse & Hol, 2017). Further study of N2O production in biofilters is
urgently required.
Figure 20 shows a conceptual model we have originated, suitable for winter cattle barns in northern Europe:
air (~100 to 200 cows, with air having 10–100 ppm CH4) is extracted from the barn via pipes laid at low level
along food troughs. In the conceptual model, the air is then drawn through adjacent greenhouse plant soils
or growth media inoculated with methanotrophs. Scheutz et al. (2009), reviewing studies of soil uptake per-
formance, showed that consumption rates of 100–400 g·m−2·day−1 or higher can potentially be achieved.
Extractor fans and light heating could be powered by solar or wind electricity and operate intermittently
when power is available. Thus, if the greenhouse is large enough (e.g., 50–100% the area of the cattle barn,
assuming typical per‐cow areas), then in principle it may be possible to remove a substantial part of the
emitted methane, reducing the methane mole fraction toward ambient amounts in the air that is finally
extracted to the atmosphere. But this attractive conceptual model hides a problem. Ammonia would be
converted to N2O, which is an important a long‐lived greenhouse gas.
The viability of such conceptual model will depend not only on demonstrated effectiveness and size of the
soil biofilter (or other material, such as gravel, woodchip or biochar) needed but also on demand (number
of greenhouses installed) and financial subsidy. Careful consideration of the biofilter energy use is also
required to ensure that the energy used to operate a filter with dilute ppm‐level emissions does not have a
higher warming potential than the emitted methane. Renewable energy sources should remove this con-
cern. In northern temperate countries, viability of methane removal by twinning greenhouses with winter
cattle barns could be enhanced by incentives favoring local greenhouse products, rather than long distance
trucking or airfreighting of winter fruit and vegetables from warmer regions.

7. Summary of Priorities
The U.S. EPA's Greenhouse Gas Equivalencies Calculator (https://www.epa.gov/energy/greenhouse‐gas‐
equivalencies‐calculator) gives a good perspective on the options for reducing climate warming impacts.
The measures advocated here to reduce methane emissions are widely varied, and mostly specific to the
sources considered. For methane mitigation, costs vary. Some are inexpensive but others will demand sub-
stantial investment or tax incentives, costly either for producers or governments. All rely on well‐understood
technology and are well proven, with the exception of some options discussed in section 6.3, such as the
application of methanotrophy to remove ruminant emissions. For a review of frontier ideas and technology,
many of which are likely to become economic to apply in the next decade, see Pratt & Tate (2018) and
Jackson et al. (2019).
Mitigating gas emissions from the fossil fuel industry is an important target. In the gas industry, better
and more comprehensive monitoring programs are needed (Fox et al., 2019). Many emissions are
deliberate vents, that can be much reduced by better design and firmer regulation. Reduction of leaks
at production sites can be driven by a mix of regulatory change, penalty fines, and small tax incentives,
with little demand on the public fiscus. In the industrialized nations, the urban gas distribution network
is leaky and antiquated. Cheap off‐peak renewable or nuclear electricity may make it possible to
abandon gas entirely for domestic purposes, and repurpose the pipeline networks for telecommunica-
tions or power delivery.
Cutting landfill emissions is a priority. In Europe there have been sharp reductions some years ago, but it is
possible the progress is stalling. In the United States and China, there is much to be done. In the tropics, the
work is barely starting. Uncontrolled emissions from huge landfills in cities like Delhi (Ghosh et al., 2019)
not only damage the quality of life of local people but hurt the environment worldwide. The prime respon-
sibility is with local jurisdictions, which should share in the burden of maintaining planetary health.
Mitigating methane emissions from agriculture (Smith et al., 2008) is an immense topic. The most obvious
target is to cut emissions from manure tanks and lagoons around large cattle feed lots and barns. These
sources are easily located and aerated, to limit anaerobic methanogenesis. More generally, there is a need
for better comparative studies of grass‐fed versus stall‐fed cattle. Intuitively, grass‐fed cattle with manure
falling in open fields should be less methane‐emitting than industrial cows (Figure 21).

NISBET ET AL. 38 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Cutting agricultural ruminant populations by changing human diets reduces methane emissions (Poore &
Nemecek, 2018). There is enough food for a substantial reduction if food is better shared and wasted less
(Campbell et al., 2017). The broader impacts of going further are less clear. Cattle are essential to India,
the world's ruminant super power, and in moist tropical Africa. The trade‐offs between providing human
food from local ruminants compared to importing arable crops are complex, even if cultural factors are
ignored. Pasture lands are usually land that is too dry, too hilly, too stony, or too infertile for crops.
Terminating production from the 22% of the planet's ice‐free land surface that is pasture (Ramankutty et
al., 2008), and replacing it with fully vegan (nonanimal) food would require more fertilizer‐intensive farming
of existing arable land and likely expansion of crop production into the remaining ploughable forest land in
the moist tropics.
Such questions as yet remain open. Detailed whole‐chain studies are needed that fully account for all
impacts, both direct (e.g., greenhouse gas emissions, including CO2 and N2O) and indirect (biodiversity
impact, etc.). Careful whole‐impact thought is essential before advocating abandoning food production from
grassland pastures and replacing it by intensive crop farming elsewhere.

8. Outlook
What will be the impact of successful mitigation? Is it worth even attempting?
The recent unexpected strong growth in the atmospheric methane burden challenges the feasibility of
meeting the 1.5 °C target of the UNFCCC Paris Agreement (Nisbet et al., 2019). Collins et al. (2018)
showed that, if the Paris target is to be met, there is a simple relationship between methane in the
Year 2100 and “allowable” carbon emissions to that year: −0.27 ± 0.05 GtC per ppb methane. Thus,
cutting methane sharply allows more relaxed CO2 targets. Conversely, inaction on methane would
imply stronger CO2 (or other forcing agent) targets. However, Solomon et al. (2013) pointed out that
although emphasis on a short‐lived forcing agent like methane will indeed help to moderate warming,
if that emphasis comes at the expense of efforts to reduce CO2 emission, then short‐term gain is paid for
by greater long‐term warming. Thus a “two‐basket” policy is wiser—to cut both CO2 and methane
burdens. Cain et al. (2019) propose the use of CO2‐warming‐equivalent emissions (CO2‐we) to assess
how different greenhouse gases contribute to global warming. Using CO2‐we emissions allow methane
and other greenhouse gases to be brought into the concept of a carbon budget and show how much
each gas contributes to warming.
Figure 22 shows what can be achieved, showing results from an atmospheric box model using a simple CO,
CH4, and OH chemical mechanism, following Prather (1996). Inputs include a prescribed OH source, along
with CH4 and CO emissions. For a small (5%) perturbation in methane emissions, the box model gives a
feedback factor (ratio of fractional change in concentration to fractional change in emissions) of ~1.5,
agreeing well with values derived from other studies, in the range 1.3 to 1.7 (Holmes et al., 2013;
Prather et al., 2001). Box model results for different possible methane emission reduction paths suggest that
emission reductions of ~25 % are required to bring global methane mixing ratios back in line with RCP 2.6
projections by 2100.
Figure 22 implies that a cut of about 30% in total methane emissions—for example, 150 Tg/a or more—is
needed to bring methane back to a pathway comparable to RCP2.6. To put this into context, the finding
by Shindell et al. (2017), that as much as 110 Tg CH4 per year of abatement is possible by scaling up existing
technology and policy options, offers an attractive route. Note, however, that RCP2.6 is only illustrative, as
one of many possible paths to the Paris Agreement goals. But if methane reduction does not happen, in order
to achieve the Paris Agreement goals, other greenhouse gases will have to be reduced more substantially in
place of methane. Are these large methane reductions feasible? They should be, if very substantial cuts in gas
and coal industry losses and landfill outputs are partnered with significant reductions in agricultural emis-
sions. Methane mitigation has no magic bullet: the sources are too diverse.
Science can only inform, not choose. Successful mitigation will require implementation of a wide array
of differing approaches, each appropriate to specific circumstances and source types. Wise policy‐making
depends on judging the balance of cost and benefit: Given the overall international reluctance of demo-
cratic taxpayers to attack emissions, what is the most acceptable compromise? Information is vital—

NISBET ET AL. 39 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

would the 2018 suspension of methane rules by the U.S. EPA (EPA, 2018; Reuters, 2018a, 2018b) have
happened if public information had been better? Given the existence of many “no regrets” options for
methane mitigation, for example, eliminating leakage of methane that can otherwise be used, or con-
suming fewer products from intensive industrial‐scale animal husbandry, it would be foolish not to
act, but there is a danger in acting without considering full whole‐impact consequences.
The challenge is considerable: Where should resources be allocated? How should global warming potential
be assessed? How can local, regional, and global impacts be balanced? If the removal of methane demands
energy, as is likely, it is also likely that the energy will have been generated by a process that emits CO2. Will
the mitigation of methane increase CO2?
Striking a balance between methane reduction and CO2 reduction involves political judgment. The
evidence presented in this review suggests the methane burden could be reduced significantly by an
investment in mitigation that is comparatively modest in comparison to the costs of parallel and neces-
sary measures to reduce CO2, and much less than the cost of climate change. Advances in detection,
characterization and quantification of methane emissions mean that methane mitigation is becoming
much easier, which potentially may tilt the near‐term balance of priority in favor of methane reduction.
The task is that of Sisyphus, who pushed a boulder up a slope only to find it rolled back just as quickly. It is
necessary to support the growing global human population while reducing the downward risk of breakdown
in the natural atmospheric maintenance system. Yet, though Sisyphus made no progress in pushing the
boulder up the hill, a better hero, Hercules, did succeed in a tougher task—he cleared the Augean stables
by adopting new technology.
Mitigating methane will need a global effort. It is not only Europe, North America, Australia, and East Asia
that need to act. Commitment is needed also from Asian and African tropical nations. Fire management and
Acknowledgmens
We thank the U.K. Natural covering landfills are obvious immediate actions that will bring rapid air quality improvement as well as
Environment Research Council for helping the global greenhouse. Burden sharing is important. International meetings on greenhouse gas
current Grants NE/P019641/1 New
reduction often focus on emissions from developed nations, but there are many actions, many at compara-
methodologies for removal of methane,
NE/N016238/1 MOYA The Global tively low cost (such as landfill cover) and with strong local benefits (such as fire reduction), that can be
Methane Budget 2016–2020, implemented by tropical nations.
NE/S004211/1 Detection and
Attribution of Regional greenhouse gas The greatest long‐term driver of change is human population growth and food needs, particularly in the tro-
Emissions in the United Kingdom pics. Diets can change. Population growth slows wherever nations provide better education, especially
(DARE‐UK), NE/R01809X1/1 (Co‐I,
BGS‐led) EQUIP4Risk Evaluation of
female education, and improved living standards. Those factors are outside the scope of this discussion.
shale gas risk, NE/S00159X/1 ZWAMPS Nevertheless, within the narrower focus of the world in 2020, application of the technical options discussed
Quanitfying methane emissions in here could indeed slow methane growth, and soon. But if methane is not brought under control, the conse-
remote tropical settings: a new 3D
approach, and for earlier NERC grants;
quences are severe. The warming may feed the warming (Dean et al., 2018). Methane mitigation is important
the UK Government's Department for and not enormously costly. It is possible, and it can be done quickly. If the Paris Agreement is to succeed,
Business, Energy and Industrial methane needs attention.
Strategy GA/18F/017/NEE6617R, the
European Union for supporting
MEMO2 (MEthane goes MObile:
MEasurement and MOdelling) and for
earlier GEOMON and other support;
References
the United Nations' Climate & Clean Abushammala, M., Basri, N. E. A., Irwan, D., & Younes, M. K. (2014). Methane oxidation in landfill cover soils: A review. Asian Journal of
Air Coalition for its support for Atmospheric Environment, 8(1), 1–14. https://doi.org/10.5572/ajae.2014.8.1.001
Methane Science Studies, the Kuwait ADBA (2018). Biogas plants in the UK‐2018. Anaerobic Digestion and Bioresources Association. https://www.shawrenewables.co.uk/
Foundation for the Advancement of biogas-plants-in-the-uk-2018/
Sciences (KFAS) Project P115‐64SC‐01, Allen, D. (2016). Attributing atmospheric methane to anthropogenic emission sources. Accounts of Chemical Research, 49(7), 1344–1350.
and the Australian Government's https://doi.org/10.1021/acs.accounts.6b00081
Cotton Research and Development Allen, G., Hollingsworth, P., Kabbabe, K., Pitt, J. R., Mead, M. I., Illingworth, S., et al. (2019). The development and trial of an unmanned
Corporation for fieldwork in the Surat aerial system for the measurement of methane flux from landfill and greenhouse gas emission hotspots. Waste Management, 87,
Basin, Australia. We particularly thank 883–892. https://doi.org/10.1016/j.wasman.2017.12.024
the US NOAA Carbon Cycle Allen, G., Illingworth, S. M., O'Shea, S. J., Newman, S., Vance, A., Bauguitte, S. J. B., et al. (2014). Atmospheric composition and ther-
Greenhouse Gases group in Boulder for modynamic retrievals from the ARIES airborne TIR‐FTS system—Part 2: Validation and results from aircraft campaigns. Atmospheric
their generous and deep help and sup- Measurement Techniques Discussions, 7(12), 4401–4416. https://doi.org/10.5194/amt-7-4401-2014
port, and the two anonymous and Allen, G., Williams, P., Shah, A., Hollingsworth, P., Kabbabe, K., Helmore, J., et al. (2018). Validation of landfill methane measurements
extremely diligent referees to whom we from an unmanned aerial system. Environment Agency. Annual Report and Accounts, [SC160006]
are very grateful for their very detailed al‐Shalaan, A. (2019). Methane emissions in Kuwait: Plume identification, isotopic characterisation & inventory verification. Unpubl. Ph.
and constructive comments throughout D. thesis, Royal Holloway, Univ. of London.
this long manuscript. No data are pre- Alvarez, R., Zavala‐Araiza, D., Lyon, D. R., Allen, D. T., Barkley, Z. R., Brandt, A. R., et al. (2018). Assessment of methane emissions from
sented in this manuscript. the U.S. oil and gas supply chain. Science, 361(6398), 186–188. https://doi.org/10.1126/science.aar7204

NISBET ET AL. 40 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Amon, T., Amon, B., Kryvoruchko, V., Zollitsch, W., Mayer, K., & Gruber, L. (2007). Biogas production from maize and dairy cattle manure
—influence of biomass composition on the methane yield. Agriculture, Ecosystems & Environment, 118(1‐4), 173–182. https://doi.org/
10.1016/j.agee.2006.05.007
Andela, N., Morton, D. C., Giglio, L., Chen, Y., van der Werf, G. R., Kasibhatla, P. S., et al. (2017). A human‐driven decline in global burned
area. Science, 356(6345), 1356–1362. https://doi.org/10.1126/science.aal4108
Andela, N., Morton, D. C., Giglio, L., Paugam, R., Chen, Y., Hantson, S., et al. (2019). The Global Fire Atlas of individual fire size, duration,
speed and direction. Earth System Science Data, 11(2), 529–552. https://doi.org/10.5194/essd-11-529-2019
Andela, N., & van der Werf, G. R. (2014). Recent trends in African fires driven by cropland expansion and El Niño to La Niña transition.
Nature Climate Change, 4(9), 791–795. https://doi.org/10.1038/nclimate2313
Andersen, T., Scheeren, B., Peters, W., & Chen, H. (2018). A UAV‐based active AirCore system for measurements of greenhouse gases.
Atmospheric Measurement Techniques, 11(5), 2683–2699. https://doi.org/10.5194/amt-11-2683-2018
Angelidaki, I., Treu, L., Tsapekos, P., Luo, G., Campanaro, S., Wenzel, H., & Kougias, P. (2018). Biogas upgrading and utilization: Current
status and perspectives. Biotechnology Advances, 36(2), 452–466. https://doi.org/10.1016/j.biotechadv.2018.01.011
Ars, S., Broquet, G., Yver Kwok, C., Roustan, Y., Wu, L., Arzoumanian, E., & Bousquet, P. (2017). Statistical atmospheric inversion of local
gas emissions by coupling the tracer release technique and local‐scale transport modelling: A test case with controlled methane emis-
sions. Atmospheric Measurement Techniques, 10(12), 5017–5037. https://doi.org/10.5194/amt-10-5017-2017
Australia's National Greenhouse Gas Accounts (2017). https://ageis.climatechange.gov.au
Baer, D., Gupta, M., Leen, J. B., & Berman, E. (2012). Environmental and atmospheric monitoring using off‐axis integrated cavity output
spectroscopy (OA‐ICOS). American Laboratory, 44, 20–23.
Baillie, J., Risk, D., Atherton, E., O'Connell, E., Fougère, C., Bourlon, E., & MacKay, K. (2019). Methane emissions from conventional and
unconventional oil and gas production sites in southeastern Saskatchewan, Canada. Environmental Research Communications, 1,
å011003.
Ball, B. C., Dobbie, K. E., Parker, J. P., & Smith, K. A. (1997). The influence of gas transport and porosity on methane oxidation in soils.
Journal of Geophysical Research, 102(D19), 23,301–23,308. https://doi.org/10.1029/97JD00870
Bamberger, I., Stieger, J., Buchmann, N., & Eugster, W. (2014). Spatial variability of methane: Attributing atmospheric concentrations to
emissions. Environmental Pollution, 190, 65–74. https://doi.org/10.1016/j.envpol.2014.03.028
Bandara, W. M. K. R. T. W., Kindaichi, T., Satoh, H., Sasakawa, M., Nakahara, Y., Takahashi, M., & Okabe, S. (2012). Anaearobic treatment
of municipal wastewater at ambient temperature: Analysis of archaeal community structure and recovery of dissolved methane. Water
Research, 46(17), 5756–5764. https://doi.org/10.1016/j.watres.2012.07.061
Barchyn, T. E., Hugenholtz, C. H., Myshak, S., & Bauer, J. (2017). A UAV‐based system for detecting natural gas leaks. Journal of
Unmanned Vehicle Systems, 6, 18–30.
Barkley, Z. R., Lauvaux, T., Davis, K. J., Deng, A., Fried, A., Weibring, P., et al. (2019). Estimating methane emissions from underground
coal and natural gas production in southwestern Pennsylvania. Geophysical Research Letters, 46, 4531–4540. https://doi.org/10.1029/
2019GL082131
Bell, C. S., Vaughn, T. L., Zimmerle, D., Herndon, S. C., Yacovitch, T. I., Heath, G. A., et al. (2017). Comparison of methane emission
estimates from multiple measurement techniques at natural gas production pads. Elementa, 5. NREL/JA‐6A20‐70835
Benchaar, C., Pomar, C., & Chiquette, J. (2001). Evaluation of dietary strategies to reduce methane production in ruminants: A modelling
approach. Canadian Journal of Animal Science, 81(4), 563–574. https://doi.org/10.4141/A00-119
Bergamaschi, P., Houweling, S., Segers, A., Krol, M., Frankenberg, C., Scheepmaker, R. A., et al. (2013). Atmospheric CH4 in the first
decade of the 21st century: Inverse modeling analysis using SCIAMACHY satellite retrievals and NOAA surface measurements. Journal
of Geophysical Research: Atmospheres, 118, 7350–7369. https://doi.org/10.1002/jgrd.50480
Bergamaschi, P., Krol, M., Meirink, J. F., Dentener, F., Segers, A., van Aardenne, J., et al. (2010). Inverse modeling of European CH4
emissions 2001‐2006. Journal of Geophysical Research, 115, D22309. https://doi.org/10.1029/2010JD014180
Bergamaschi, P., Karstens, U., Manning, A. J., Saunois, M., Tsuruta, A., Berchet, A., et al. (2018). Inverse modelling of European CH 4
emissions during 2006–2012 using different inverse models and reassessed atmospheric observations. Atmospheric Chemistry and
Physics, 18(2), 901–920. https://doi.org/10.5194/acp-18-901-2018
Boeckx, P., Cleemput, O., & Villaralvo, I. (1996). Methane emission from a landfill and the methane oxidising capacity of its covering soil.
Soil Biology and Biochemistry, 28(10‐11), 1397–1405. https://doi.org/10.1016/S0038-0717(96)00147-2
Boucher, O., & Folberth, G. A. (2010). New directions: Atmospheric methane removal as a way to mitigate climate change? Atmospheric
Environment, 44(27), 3343–3345. https://doi.org/10.1016/j.atmosenv.2010.04.032
Brandt, A. R., Heath, G. A., Kort, E. A., O'Sullivan, F., Petron, G., Jordaan, S. M., et al. (2014). Methane leaks from North American natural
gas systems. Science, 343(6172), 733–735. https://doi.org/10.1126/science.1247045
Brownlow, R., Lowry, D., Fisher, R. E., France, J. L., Lanoisellé, M., White, B., et al. (2017). Isotopic ratios of tropical methane emissions by
atmospheric measurement. Global Biogeochemical Cycles, 31, 1408–1419. https://doi.org/10.1002/2017GB005689
Brownlow, R., Lowry, D., Thomas, R. M., Fisher, R. E., France, J. L., Cain, M., et al. (2016). Methane mole fraction and δ13C above and
below the trade wind inversion at Ascension Island in air sampled by aerial robotics. Geophysical Research Letters, 43, 11,893–11,902.
https://doi.org/10.1002/2016GL071155
Buciuman, F. C., Patcas, F., & Hahn, T. (1999). A spillover approach to oxidation catalysis over copper and manganese mixed oxides.
Chemical Engineering and Processing, 38(4‐6), 563–569. https://doi.org/10.1016/S0255-2701(99)00053-7
Buddle, B. M., Denis, M., Attwood, G. T., Altermann, E., Janssen, P. H., Ronimus, R. S., et al. (2011). Strategies to reduce methane emissions
from farmed ruminants grazing on pasture. The Veterinary Journal, 188(1), 11–17. https://doi.org/10.1016/j.tvjl.2010.02.019
Cain, M., Lynch, J., Allen, M. R., Fuglestvedt, J. S., Frame, D. J., & Macey, A. H. (2019). Improved calculation of warming‐equivalent
emissions for short‐lived climate pollutants. npj Climate and Atmospheric Science, 2(1), 29. https://doi.org/10.1038/s41612-019-0086-4
Cain, M., Warwick, N. J., Fisher, R. E., Lowry, D., Lanoisellé, M., Nisbet, E. G., et al. (2017). A cautionary tale: A study of a methane
enhancement over the North Sea. Journal of Geophysical Research: Atmospheres, 122, 7630–7645. https://doi.org/10.1002/2017JD026626
Cambaliza, M. O. L., Shepson, P. B., Bogner, J., Caulton, D. R., Stirm, B., Sweeney, C., et al. (2015). Quantification and source apportion-
ment of the methane emission flux from the city of Indianapolis. Elementa: Science of the Anthropocene, 3, 000037. https://doi.org/
10.12952/journal.elementa.000037
Campbell, B. M., Beare, D. J., Bennett, E. M., Hall‐Spencer, J. M., Ingram, J. S., Jaramillo, F., et al. (2017). Agriculture production as a major
driver of the Earth system exceeding planetary boundaries. Ecology and Society, 22, 8. https://doi.org/10.5751/ES-09595-220408
Canada (2016). Canada‐U.S. Joint Statement on Climate, Energy and Arctic. https://pm.gc.ca/eng/news/2016/03/10/us-canada-joint-
statement-climate-energy-and-arctic-leadership

NISBET ET AL. 41 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Caulton, D. R., Li, Q., Bou‐Zeid, E., Fitts, J. P., Golston, L. M., Pan, D., et al. (2018). Quantifying uncertainties from mobile‐laboratory‐
derived emissions of well pads using inverse Gaussian methods. Atmospheric Chemistry and Physics, 18(20), 15,145–15,168. https://doi.
org/10.5194/acp-18-15145-2018
Caulton, D. R., Shepson, P. B., Santoro, R. L., Sparks, J. P., Howarth, R. W., Ingraffea, A. R., et al. (2014). Toward a better understanding and
quantification of methane emissions from shale gas development. Proceedings of the National Academy of Sciences, 111(17), 6237–6242.
https://doi.org/10.1073/pnas.1316546111
Chang, J., Peng, S., Ciais, P., Saunois, M., Dangal, S. R. S., Herrero, M., et al. (2019). Revisiting enteric methane emissions from domestic
ruminants and their δ13CCH4 source signature. Nature Communications, 10(1), 3420. https://doi.org/10.1038/s41467-019-11066-3
Chanton, J., Abichou, T., Langford, C., Hater, G., Green, R., Goldsmith, D., & Swan, N. (2011). Landfill methane oxidation across climate
types in the US. Environmental Science & Technology, 45(1), 313–319. https://doi.org/10.1021/es101915r
Chen, Y., Morton, D. C., Andela, N., Van Der Werf, G. R., Giglio, L., & Randerson, J. T. (2017). A pan‐tropical cascade of fire driven by El
Niño/Southern Oscillation. Nature Climate Change, 7(12), 906–911. https://doi.org/10.1038/s41558-017-0014-8
Chen, J., Dietrich, F., Maazallahi, H., Forstmaier, A., Winkler, D., Hofmann, M. E., et al. (2019). Methane Emissions from the Munich
Oktoberfest. Atmospheric Chemistry and Physics Discussion. https://doi.org/10.5194/acp-2019-709
Christiansen, J., Levy‐Booth, D., Prescott, C. E., & Grayston, S. J. (2016). Different soil moisture control of net methane oxidation and
production in organic upland and wet forest soils of the Pacific coastal rainforest in Canada. Canadian Journal of Forest Research, 47,
628–635.
Collier‐Oxandale, A., Casey, J. G., Piedrahita, R., Ortega, J., Halliday, H., Johnston, J., & Hannigan, M. P. (2018). Assessing a low‐cost
methane sensor quantification system for use in complex rural and urban environments. Atmospheric Measurement Techniques, 11(6),
3569–3594. https://doi.org/10.5194/amt-11-3569-2018
Collins, M., Knutti, R., Arblaster, J., Dufresne, J. L., Fichefet, T., & Friedlingstein, P. (2013). Section 12.3.1.3 The new concentration driven
RCP scenarios, and their extensions. In T. F. Stocker, et al. (Eds.), Climate Change 2013: The Physical Science Basis. Contribution of
Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Chap. 12, pp. 1029–1136). Cambridge,
UK and New York, USA: Cambridge University Press.
Collins, W. J., Webber, C. P., Cox, P. M., Huntingford, C., Lowe, J., Sitch, S., et al. (2018). Increased importance of methane reduction for a
1.5 degree target. Environmental Research Letters, 13(5), 054003. https://doi.org/10.1088/1748-9326/aab89c
Committee on Acute Exposure Guideline Levels (2013. Ethyl Mercaptan Acute Exposure Guideline Levels). Committee on toxicology; Board
on Environmental Studies and Toxicology; Division on Earth and Life Studies; National Reserch Council. Acute Exposure Guideline Levels
for Selected Airborne Chemicals (Vol. 15). Washington (DC): Available from:National Academies Press (US). 2013 Sep 26. 1https://www.
ncbi.nlm.nih.gov/books/NBK201325/
Conley, S., Franco, G., Faloona, I., Blake, D. R., Peischl, J., & Ryerson, T. B. (2016). Methane emissions from the 2015 Aliso Canyon blowout
in Los Angeles, CA. Science, 351(6279), 1317–1320. https://doi.org/10.1126/science.aaf2348
Costa, A., Chiarello, G. L., Selli, E., & Guarino, M. (2012). Effects of TiO2 based photocatalytic paint on concentrations and emissions of
pollutants and on animal performance in a swine weaning unit. Journal of Environmental Management, 96(1), 86–90. https://doi.org/
10.1016/j.jenvman.2011.08.025
Crosson, E. (2008). A cavity ring‐down analyzer for measuring atmospheric levels of methane, carbon dioxide, and water vapor. Applied
Physics B, 92(3), 403–408. https://doi.org/10.1007/s00340-008-3135-y
Cui, Y. Y., Henze, D. K., Brioude, J., Angevine, W. M., Liu, Z., Bousserez, N., et al. (2019). Inversion estimates of lognormally distributed
methane emission rates from the Haynesville‐Bossier oil and gas production region using airborne measurements. Journal of
Geophysical Research: Atmospheres, 124, 3520–3531. https://doi.org/10.1029/2018JD029489
Cusworth, D. H., Jacob, D. J., Sheng, J. X., Benmergui, J., Turner, A. J., Brandman, J., et al. (2018). Detecting high‐emitting methane
sources in oil/gas fields using satellite observations. Atmospheric Chemistry and Physics, 18(23), 16,885–16,896.
Dairy Coordinated Agricultural Project (2019). Univ. of Wisconsin & US Department of Agriculture. https://uwmadison.app.box.com/s/
q7pqa9zi2zsh8tg2e5iz4tlhu2zfh6te
Danby, G. (1998) Regulating gas utilities. House of Commons Library Research Paper 98/19 researchbriefings.files.parliament.uk/docu-
ments/RP98-19/RP98-19.pdf
Day, S., Ong, C., Rodger, A., Etheridge, D., Hibberd, M., van Gorsel Spencer, D., et al. (2015) Characterisation of regional fluxes of methane
in the Surat Basin, Queensland: Phase 2: A pilot study of methodology to detect and quantify methane sources. CSIRO, Australia. Report
for the Gas Industry Social and Environmental Research Alliance (GISERA), Project No GAS1315, EP 15369 https://gisera.csiro.au/wp-
content/uploads/2018/03/GHG-1-Phase-2-Report.pdf
de Richter, R., Tingzhen, M., Davies, P., Wei, L., & Caillol, S. (2017). Removal of non‐CO2 greenhouse gases by large‐scale atmospheric
solar photocatalysis. Progress in Energy and Combustion Science, 60, 68–96.
Dean, J. F., Middelburg, J. J., Röckmann, T., Aerts, R., Blauw, L. G., Egger, M., et al. (2018). Methane feedbacks to the global climate system
in a warmer world. Reviews of Geophysics, 56, 207–250. https://doi.org/10.1002/2017RG000559
Deetz, K., & Vogel, B. (2017). Development of a new gas-flaring emission dataset for southern West Africa. Geoscientific Model Development,
10, 1607–1620.
Detto, M., Verfaillie, J., Anderson, F., Xu, L., & Baldocchi, D. (2011). Comparing laser‐based open‐and closed‐path gas analyzers to measure
methane fluxes using the eddy covariance method. Agricultural and Forest Meteorology, 151(10), 1312–1324. https://doi.org/10.1016/j.
agrformet.2011.05.014
Dlugokencky, E. J., Nisbet, E. G., Fisher, R. E., & Lowry, D. (2011). Global atmospheric methane: Budget changes, and dangers.
Philosophical Transactions of the Royal Society A, 369(1943), 2058–2072. https://doi.org/10.1098/rsta.2010.0341
Döbber, D., Kießling, D., Schmitz, W., & Wendt, G. (2004). MnOx/ZrO2 catalysts for the total oxidation of methane and chloromethane.
Applied Catalysis B: Environmental, 52(2), 135–143. https://doi.org/10.1016/j.apcatb.2004.02.012
Domenikos, S. P., Vogel, F. R., Murphy, J. G., Moran, M. D., Stroud, C. A., Ren, S., et al. (2019). Towards understanding the variability in
source contribution of CO2 using high‐resolution simulations of atmospheric δ13CO2 signatures in the Greater Toronto Area, Canada.
Atmospheric Environment, 214, 116877. https://doi.org/10.1016/j.atmosenv.2019.116877
Dontala, S. P., Reddy, T. B., & Vadde, R. (2015). Environmental aspects and impacts its mitigation measures of corporate coal mining.
Procedia Earth and Planetary Science, 11, 2–7. https://doi.org/10.1016/j.proeps.2015.06.002
Dunmore, R., Hopkins, J., Lidster, R., Mead, M., Bandy, B., Forster, G., et al. (2016). Development of a combined heart‐cut and compre-
hensive two‐dimensional gas chromatography system to extend the carbon range of volatile organic compounds analysis in a single
instrument. Separations, 3(3), 21. https://doi.org/10.3390/separations3030021

NISBET ET AL. 42 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Duren, R. M., Thorpe, A. K., Foster, K. T., Rafiq, T., Hopkins, F. M., Yadav, V., et al. (2019). California's methane super‐emitters. Nature,
575(7781), 180–184. https://doi.org/10.1038/s41586-019-1720-3
EDGAR‐European Commission, Joint Research Center/Netherlands Environmental Assessment Agency (2011). Emission Database for
Global Atmospheric Research (EDGAR) (version 4.2); http://edgar.jrc.ec.europa.eu.
Einola, J.‐K. M., Kettunen, R. H., & Rintala, J. A. (2007). Responses of methane oxidation to temperature and water content in cover soil of a
boreal landfill. Soil Biology and Biochemistry, 39(5), 1156–1164. https://doi.org/10.1016/j.soilbio.2006.12.022
Elvidge, C., Ziskin, D., Baugh, K., Tuttle, B., Ghosh, T., Pack, D., Erwin, E., & Zhizhin, M. (2009). A fifteen year record of global natural gas
flaring derived from satellite data. Energies, 2, 595–622.
Elvidge, C., Zhizhin, M., Baugh, K., Hsu, F. C., & Ghosh, T. (2016). Methods for global survey of natural gas flaring from visible infrared
imaging radiometer suite data. Energies, 9, 14.
Emran, B. J., Tannant, D. D., & Najjaran, H. (2017). Low‐altitude aerial methane concentration mapping. Remote Sensing, 9(8), 823. https://
doi.org/10.3390/rs9080823
Englander, J. G., Brandt, A. R., Conley, S., Lyon, D. R., & Jackson, R. B. (2018). Aerial interyear comparison and quantification of methane
emissions persistence in the Bakken Formation of North Dakota, USA. Environmental Science & Technology, 52(15), 8947–8953. https://
doi.org/10.1021/acs.est.8b01665
EPA (2018). Actions and notices about oil and natural gas air pollution standards. US Environmental Protection Agency. https://www.epa.
gov/controlling-air-pollution-oil-and-natural-gas-industry/actions-and-notices-about-oil-and-natural-gas
Etminan, M., Myhre, G., Highwood, E. J., & Shine, K. P. (2016). Radiative forcing of carbon dioxide, methane and nitrous oxide: A sig-
nificant revision of the methane radiative forcing. Geophysical Research Letters, 43, 12,614–12,623. https://doi.org/10.1002/
2016GL071930
EBA (2019). Gas for climate: extended analysis on the optimal role for gas in a net zero emissions energy system. European Biogas
Association. https://www.europeanbiogas.eu/the-optimal-role-for-gas-in-a-netzero-emissions-energy-system/
Feinberg, A. I., Coulon, A., Stenke, A., Schwietzke, S., & Peter, T. (2018). Isotopic source signatures: Impact of regional variability
on the d13CH4 trend and spatial distribution. Atmsopheric Environment, 174, 99–111. https://doi.org/10.1016/j.
atmosenv.2017.11.037
Fellmann, T., Witzke, P., Weiss, F., Van Doorslaer, B., Drabik, D., Huck, I., et al. (2018). Major challenges of integrating agriculture into
climate change mitigation policy frameworks. Mitigation and Adaptation Strategies for Global Change, 23(3), 451–468. https://doi.org/
10.1007/s11027-017-9743-2
Fernández, J., Marín, P., Díez, F. V., & Ordóñez, S. (2016). Combustion of coal mine ventilation air methane in a regenerative combustor
with integrated adsorption: Reactor design and optimization. Applied Thermal Engineering, 102(2016), 167–175. https://doi.org/10.1016/
j.applthermaleng.2016.03.171
Fisher, R., Lowry, D., Wilkin, O., Sriskantharajah, S., & Nisbet, E. G. (2006). High-precision, automated stable isotope analysis of atmo-
spheric methane and carbon dioxide using continuous‐flow isotope‐ratio mass spectrometry. Rapid Communications in Mass
Spectrometry 20, 200–208.
Flesch, T. K., Desjardins, R. L., & Worth, D. (2011). Fugitive methane emissions from an agricultural biodigester. Biomass and Bioenergy,
35(9), 3927–3935. https://doi.org/10.1016/j.biombioe.2011.06.009
Forkel, M., Dorigo, W.A., Lasslop, G., Chuvieco, E., Hantson, S., Heil, A., et al. (2019). Recent global and regional trends in burned area and
their compensating environmental controls. Environmental Research Communications.
Foster‐Wittig, T. A., Thoma, E. D., Green, R. B., Hater, G. R., Swan, N. D., & Chanton, J. P. (2015). Development of a mobile tracer cor-
relation method for assessment of air emissions from landfills and other area sources. Atmospheric Environment, 102, 323–330. https://
doi.org/10.1016/j.atmosenv.2014.12.008
Fox, T. A., Barchyn, T. E., Risk, D., Ravikumar, A. P., & Hugenholtz, C. H. (2019). A review of close‐range and screening technologies for
mitigating fugitive methane emissions in upstream oil and gas. Environmental Research Letters, 14(5), 053002. https://doi.org/10.1088/
1748-9326/ab0cc3
Frankenberg, C., Meirink, J. F., Bergamaschi, P., Goede, A. P. H., Heimann, M., Körner, S., et al. (2006). Satellite chartography of atmo-
spheric methane from SCIAMACHY on board ENVISAT: Analysis of the years 2003 and 2004. Journal of Geophysical Research, 111,
D07303. https://doi.org/10.1029/2005JD006235
Fredenslund, A., Hinge, J., Holmgren, M., Rasmussen, S., & Scheutz, C. (2018). On‐site and ground‐based remote sensing measurements of
methane emissions from four biogas plants: A comparison study. Bioresource Technology, 270(June), 88–95. https://doi.org/10.1016/j.
biortech.2018.08.080
Fries, A. E., Schifman, L. A., Shuster, W. D., & Townsend‐Small, A. (2018). Street‐level emissions of methane and nitrous oxide from
the wastewater collection system in Cincinnati, Ohio. Environmental Pollution, 236, 247–256. https://doi.org/10.1016/j.
envpol.2018.01.076
Gebert, J., & Gröngröft, A. (2006). Performance of a passively vented field‐scale biofilter for the microbial oxidation of landfill methane.
Waste Management, 26(4), 399–407. https://doi.org/10.1016/j.wasman.2005.11.007
Ghosh, P., Shah, G., Chandra, R., Sahota, S., Kumar, H., Vijay, V. K., & Thakur, I. S. (2019). Assessment of methane emissions and energy
recovery potential from the municipal solid waste landfills of Delhi, India. Bioresource Technology, 272, 611–615. https://doi.org/
10.1016/j.biortech.2018.10.069
Gilbert, M., Nicolas, G., Cinardi, G., Van Boeckel, T. P., Vanwambeke, S. O., Wint, G. W., & Robinson, T. P. (2018). Global distribution data
for cattle, buffaloes, horses, sheep, goats, pigs, chickens and ducks in 2010. Scientific data, 5(1), 180227. https://doi.org/10.1038/
sdata.2018.227
Giménez, J. B., Martí, N., Ferrer, J., & Seco, A. (2012). Methane recovery efficiency in a submerged anaerobic membrane bioreactor
(SAnMBR) treating sulphate‐rich urban wastewater: Evaluation of methane losses with the effluent. Bioresource Technology, 118, 67–72.
https://doi.org/10.1016/j.biortech.2012.05.019
Gioli, B., Toscano, P., Lugato, E., Matese, A., Miglietta, F., Zaldei, A., & Vaccari, F. P. (2012). Methane and carbon dioxide fluxes and source
partitioning in urban areas: The case study of Florence, Italy. Environmental Pollution, 164, 125–131. https://doi.org/10.1016/j.
envpol.2012.01.019
Girard, M., Ramirez, A. A., Buelna, G., & Heitz, M. (2011). Biofiltration of methane at low concentrations representative of the piggery
industry—Influence of the methane and nitrogen concentrations. Chemical Engineering Journal, 168(1), 151–158. https://doi.org/
10.1016/j.cej.2010.12.054
Golston, L. M., Tao, L., Brosy, C., Schäfer, K., Wolf, B., McSpiritt, J., et al. (2017). Lightweight mid‐infrared methane sensor for unmanned
aerial systems. Applied Physics B, 123(6), 170. https://doi.org/10.1007/s00340-017-6735-6

NISBET ET AL. 43 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Gosiewski, K., & Pawlaczyk, A. (2014). Catalytic or thermal reversed flow combustion of coal mine ventilation air methane: What is better
choice and when? Chemical Engineering Journal, 238, 78–85. https://doi.org/10.1016/j.cej.2013.07.039
Graetzel, M., Thampi, K. R., & Kiwi, J. (1989). Methane oxidation at room temperature and atmospheric pressure activated by light via
polytungstate dispersed on Titania. The Journal of Physical Chemistry, 93(10), 4128–4132. https://doi.org/10.1021/j100347a050
Grainger, C., Clarke, T., McGinn, S. M., Auldist, M. J., Beauchemin, K. A., Hannah, M. C., et al. (2007). Methane emissions from dairy cows
measured using the sulfur hexafluoride (SF6) tracer and chamber techniques. Journal of Dairy Science, 90(6), 2755–2766. https://doi.org/
10.3168/jds.2006-697
Greatwood, C., Richardson, T., Freer, J., Thomas, R., MacKenzie, A., Brownlow, R., et al. (2017). Atmospheric sampling on Ascension
Island using multirotor UAVs. Sensors, 17(6), 1189. https://doi.org/10.3390/s17061189
Guarino, M., Costa, A., & Porro, M. (2008). Photocatalytic TiO2 coating—To reduce ammonia and greenhouse gases concentration and
emission from animal husbandries. Bioresource Technology, 99(7), 2650–2658. https://doi.org/10.1016/j.biortech.2007.04.025
Guisasola, A., de Haas, D., Keller, J., & Yuan, Z. (2008). Methane formation in sewer systems. Water Research, 42(6‐7), 1421–1430.
https://doi.org/10.1016/j.watres.2007.10.014
Guisasola, A., Sharma, K. R., Keller, J., & Yuan, Z. (2009). Development of a model for assessing methane formation in rising main sewers.
Water Research, 43(11), 2874–2884. https://doi.org/10.1016/j.watres.2009.03.040
Guj, P., Fallon, M., McCuaig, T. C., & Fagan, R. (2011). A time‐series audit of Zipf's law as a measure of terrane endowment and maturity in
mineral exploration. Economic Geology, 106(2), 241–259. https://doi.org/10.2113/econgeo.106.2.241
Hacker, J. M., Chen, D., Bai, M., Ewenz, C., Junkermann, W., Lieff, W., et al. (2016). Using airborne technology to quantify and apportion
emissions of CH4 and NH3 from feedlots. Animal Production Science, 56, 190–203.
Hansen, J., Sato, M., Ruedy, R., Lacis, A., & Oinas, V. (2000). Global warming in the twenty-first century: An alternative scenario.
Proceedings of the National Academy of Sciences, 97, 9875–9880.
Hartery, S., Commane, R., Lindaas, J., Sweeney, C., Henderson, J., Mountain, M., et al. (2018). Estimating regional‐scale methane flux and
budgets using CARVE aircraft measurements over Alaska. Atmospheric Chemistry and Physics, 18(1), 185–202. https://doi.org/10.5194/
acp-18-185-2018
Hausman, C., & Muehlenbachs, L., (2016). Price regulation and environmental externalities: evidence from methane leaks (No. w22261).
National Bureau of Economic Research.
Haustein, K., Allen, M. R., Forster, P. M., Otto, F. E. L., Mitchell, D. M., Matthews, H. D., & Frame, D. J. (2017). A real‐time Global
Warming Index. Scientific Reports, 7, 15417. https://doi.org/10.1038/s41598-017-14828-5
Hay, S., Obee, T., Luo, Z., Jiang, T., Meng, Y., He, J., et al. (2015). The viability of photocatalysis for air purification. Molecules, 20(1),
1319–1356. https://doi.org/10.3390/molecules20011319
He, L., Zeng, Z. C., Pongetti, T., Wong, C., Liang, J., Gurney, K. R., et al. (2019). Atmospheric methane emissions correlate with natural gas
consumption from residential and commercial sectors in Los Angeles. Geophysical Research Letters, 46, 8563–8571. https://doi.org/
10.1029/2019GRL083400
Heimburger, A. M. F., Harvey, R. M., Shepson, P. B., Stirm, B. H., Gore, C., Turnbull, J., et al. (2017). Assessing the optimized precision of
the aircraft mass balance method for measurement of urban greenhouse gas emission rates through averaging. Elementa: Science of the
Anthropocene, 5(0), 26. https://doi.org/10.1525/elementa.134
Helfter, C., Mullinger, N., Vieno, M., O'Doherty, S., Ramonet, M., Palmer, P. I., & Nemitz, E. (2019). Country‐scale greenhouse gas budgets
using shipborne measurements: A case study for the UK and Ireland. Atmospheric Chemistry and Physics, 19(5), 3043–3063. https://doi.
org/10.5194/acp-19-3043-2019
Helfter, C., Tremper, A. H., Halios, C. H., Kotthaus, S., Bjorkegren, A., Grimmond, C. S. B., et al. (2016). Spatial and temporal variability of
urban fluxes of methane, carbon monoxide and carbon dioxide above London, UK. Atmospheric Chemistry and Physics, 16(16),
10,543–10,557. https://doi.org/10.5194/acp-16-10543-2016
Hendrick, M. F., Ackley, R., Sanaie‐Movahed, B., Tang, X., & Phillips, N. G. (2016). Fugitive methane emissions from leak‐prone natural
gas distribution infrastructure in the urban environment. Environmental Pollution, 213, 710–716. https://doi.org/10.1016/j.
envpol.2016.01.094
Henne, S., Brunner, D., Oney, B., Leuenberger, M., Eugster, W., Bamberger, I., et al. (2016). Validation of the Swiss methane emission
inventory by atmospheric observations and inverse modelling. Atmospheric Chemistry and Physics, 16(6), 3683–3710, https://doi.org/
10.5194/acp-16-3683-2016
Hiller, R. V., Neininger, B., Brunner, D., Gerbig, C., Bretscher, D., Künzle, T., et al. (2014). Aircraft‐based CH4 flux estimates for validation
of emissions from an agriculturally dominated area in Switzerland. Journal of Geophysical Research: Atmospheres, 119, 4874–4887.
https://doi.org/10.1002/2013JD020918
Holmes, C. D., Prather, M. J., Søvde, O. A., & Myhre, G. (2013). Future methane, hydroxyl, and their uncertainties: Key climate and
emission parameters for future predictions. Atmospheric Chemistry and Physics, 13(1), 285–302. https://doi.org/10.5194/acp-13-285-2013
Holmes, R. I. (2016). Mitigating ventilation air cost‐effectively from a colliery in Australia. Journal of Applied Engineering Sciences, 6(202),
41–50.
Hopkins, F. M., Ehleringer, J. R., Bush, S. E., Duren, R. M., Miller, C. E., Lai, C. T., et al. (2016). Mitigation of methane emissions in cities:
How new measurements and partnerships can contribute to emissions reduction strategies. Earth's Future, 4(9), 408–425. https://doi.
org/10.1002/2016EF000381
Houweling, S., Krol, M., Bergamaschi, P., Frankenberg, C., Dlugokencky, E. J., Morino, I., et al. (2014). A multi‐year methane inversion
using SCIAMACHY, accounting for systematic errors using TCCON measurements. Atmospheric Chemistry and Physics, 14(8),
3991–4012. https://doi.org/10.5194/acp-14-3991-2014
Hu, H., Landgraf, J., Detmers, R., Borsdorff, T., aan de Brugh, J., Aben, I., et al. (2018). Toward global mapping of methane with TROPOMI:
First results and intersatellite comparison to GOSAT. Geophysical Research Letters, 45, 3682–3689. https://doi.org/10.1002/
2018GL077259
Humpenöder, F., Popp, A., Dietrich, J. P., Klein, D., Lotze‐Campen, H., Bonsch, M., et al. (2014). Investigating afforestation and bioenergy
CCS as climate change mitigation strategies. Environmental Research Letters, 9(6), 064029. https://doi.org/10.1088/1748-9326/9/6/
064029
Iacobuta, G., Dubash, N. K., Upadhyaya, P., Deribe, M., & Höhne, N. (2018). National climate change mitigation legislation, strategy and
targets: A global update. Climate Policy, 18(9), 1114–1132. https://doi.org/10.1080/14693062.2018.1489772
IPCC (2018) Global warming of 1.5°C: A special report on the impacts of global warming of 1.5°C above pre‐industrial levels. Summary for
Policymakers. Draft report, 48th IPCC session, Incheon, Korea, 6 Oct. 2018.

NISBET ET AL. 44 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Iverach, C. P., Cendón, D. I., Hankin, S. I., Lowry, D., Fisher, R. E., France, J. L., et al. (2015). Assessing connectivity between an overlying
aquifer and a coal seam gas resource using methane isotopes, dissolved organic carbon and tritium. Scientific Reports, 5(1), 15996.
https://doi.org/10.1038/srep15996
Jackson, R. B., Down, A., Phillips, N. G., Ackley, R. C., Cook, C. W., Plata, D. L., & Zhao, K. (2014). Natural gas pipeline leaks across
Washington D.C. Environmental Science & Technology, 48(3), 2051–2058. https://doi.org/10.1021/es404474x
Jackson, R. B., Solomon, E. I., Canadell, J. G., Cargnello, M., & Field, C.B. (2019). Methane removal and atmospheric restoration. Nature
Sustainability, p.1.
Jacob, D. J., Turner, A. J., Maasakkers, J. D., Sheng, J., Sun, K., Liu, X., et al. (2016). Satellite observations of atmospheric methane and their
value for quantifying methane emissions. Atmospheric Chemistry and Physics, 16(22), 14,371–14,396. https://doi.org/10.5194/acp-16-
14371-2016
Jiang, H., Chen, Y., Jiang, P., Zhang, C., Smith, T. J., Murrell, J. C., & Xing, X. H. (2010). Methanotrophs: Multifunctional bacteria with
promising applications in environmental bioengineering. Biochemical Engineering Journal, 49(3), 277–288. https://doi.org/10.1016/j.
bej.2010.01.003
Jiang, X., Mira, D., & Cluff, D. L. (2018). The combustion mitigation of methane as a non‐CO2 greenhouse gas. Progress in Energy and
Combustion Science, 66, 176–199. https://doi.org/10.1016/j.pecs.2016.06.002
Ju, Y., Sun, Y., Sa, Z., Pan, J., Wang, J., Hou, Q., et al. (2016). A new approach to estimate fugitive methane emissions from coal mining in
China. Science of the Total Environment, 543, 514–523. https://doi.org/10.1016/j.scitotenv.2015.11.024
Kang, M., Christian, S., Celia, M. A., Mauzerall, D. L., Bill, M., Miller, A. R., et al. (2016). Identification and characterization of high
methane‐emitting abandoned oil and gas wells. Proceedings of the National Academy of Sciences of the United States of America, 113(48),
13,636–13,641. https://doi.org/10.1073/pnas.1605913113
Kang, M., Kanno, C. M., Reid, M. C., Zhang, X., Mauzerall, D. L., Celia, M. A., et al. (2014). Direct measurements of methane emissions
from abandoned oil and gas wells in Pennsylvania. Proceedings of the National Academy of Sciences, 111(51), 18,173–18,177. https://doi.
org/10.1073/pnas.1408315111
Kang, M., Mauzerall, D. L., Ma, D. Z., & Celia, M. A. (2019). Reducing methane emissions from abandoned oil and gas wells: Strategies and
costs. Energy Policy, 132, 594–601. https://doi.org/10.1016/j.enpol.2019.05.045
Karacan, C. Ö. (2011). Coal mine methane: A review of capture and utilization practices with benefits to mining safety and to greenhouse
gas reduction. International Journal of Coal Geology, 86(2‐3), 121–156. https://doi.org/10.1016/j.coal.2011.02.009
Karion, A., Lauvaux, T., Lopez Coto, I., Sweeney, C., Mueller, K., Gourdji, S., et al. (2019). Intercomparison of atmospheric trace gas dis-
persion models: Barnett Shale case study. Atmospheric Chemistry and Physics, 19, 2561–2576.
Karion, A., Sweeney, C., Kort, E. A., Shepson, P. B., Brewer, A., Cambaliza, M., et al. (2015). Aircraft‐based estimate of total methane
emissions from the Barnett Shale Region Environ. Science and Technology, 49(13), 8124–8131. https://doi.org/10.1021/acs.est.5b00217
Karion, A., Sweeney, C., Pétron, G., Frost, G., Michael Hardesty, R., Kofler, J., et al. (2013). Methane emissions estimate from airborne
measurements over a western United States natural gas field. Geophysical Research Letters, 40, 4393–4397. https://doi.org/10.1002/
grl.50811
Karion, A., Sweeney, C., Tans, P., & Newberger, T. (2010). AirCore: An innovative atmospheric sampling system. Journal of Atmospheric
and Oceanic Technology, 27(11), 1839–1853. https://doi.org/10.1175/2010JTECHA1448.1
Keeling, C. D. (1958). The concentration and isotopic abundances of atmospheric carbon dioxide in rural areas. Geochimica et
Cosmochimica Acta, 13(4), 322–334. https://doi.org/10.1016/0016-7037(58)90033-4
Kelly, B. F. G., Iverach, C. P., Lowry, D., Fisher, R. E., France, J. L. & Nisbet, E. G. (2015). Fugitive methane emissions from natural, urban,
agricultural, and energy‐production landscapes of eastern Australia. Geophysical Research Abstracts Vol. 17, EGU2015‐5135.
http://meetingorganizer.copernicus.org/EGU2015/EGU2015-5135.pdf
Kille, N., Chiu, R., Frey, M., Hase, F., Sha, M. K., Blumenstock, T., et al. (2019). Separation of methane emissions from agricultural
and natural gas sources in the Colorado Front Range. Geophysical Research Letters, 46, 3990–3998. https://doi.org/10.1029/
2019GL082132
Kormi, T., Abichou, T., Kout, N., Ksibi, M., & Wang, C. (2018). Using methane biological oxidation in soil as a tool to finance closure of
dumpsites across the Mediterranean Basin. Euro‐Mediterranean Journal for Environmental Integration, 3(1), 6. https://doi.org/10.1007/
s41207-017-0044-7
Kort, E. A., Frankenberg, C., Costigan, K. R., Lindenmaier, R., Dubey, M. K., & Wunch, D. (2014). Four corners: The largest US methane
anomaly viewed from space. Geophysical Research Letters, 41, 6898–6903. https://doi.org/10.1002/2014GL061503
Krüger, M., Beckmann, S., Engelen, B., Thielemann, T., Cramer, B., Schippers, A., & Cypionka, H. (2008). Microbial methane formation
from hard coal and timber in an abandoned coal mine. Geomicrobiology Journal, 6, 315–321.
Kumar, S., Gaikwad, S. A., Shekdar, A. V., Kshirsagar, P. S., & Singh, R. N. (2004). Estimation method for national methane emission from
solid waste landfills. Atmospheric Environment, 38(21), 3481–3487. https://doi.org/10.1016/j.atmosenv.2004.02.057
Kuze, A., Suto, H., Nakajima, M., & Hamazaki, T. (2009). Thermal and near infrared sensor for carbon observation Fourier‐transform
spectrometer on the Greenhouse Gases Observing Satellite for greenhouse gases monitoring. Applied Optics, 48(35), 6716–6733.
https://doi.org/10.1364/AO.48.006716
Kvist, T., & Aryal, N. (2019). Methane loss from commercially operating biogas upgrading plants. Waste Management, 87, 295–300.
https://doi.org/10.1016/j.wasman.2019.02.023
Lamb, B. K., Edburg, S. L., Ferrara, T. W., Howard, T., Harrison, M. R., Kolb, C. E., et al. (2015). Direct measurements show decreasing
methane emissions from natural gas local distribution systems in the United States. Environmental Science & Technology, 49(8),
5161–5169. https://doi.org/10.1021/es505116p
Lamb, B. K., McManus, J. B., Shorter, J. H., Kolb, C. E., Mosher, B., Harriss, R. C., et al. (1995). Development of atmospheric tracer methods
to measure methane emissions from natural gas facilities and urban areas. Environmental Science & Technology, 29(6), 1468–1479.
https://doi.org/10.1021/es00006a007
Lan, X., Tans, P., Sweeney, C., Andrews, A., Dlugokencky, E., Schwietzke, S., et al. (2019). Long‐term measurements show little evidence
for large increases in total U.S. methane emissions over the past decade. Geophysical Research Letters, 46, 4991–4999, https://doi.org/
10.1029/2018GL081731
Lange, A. D., & Landgraf, J. (2018). Methane profiles from GOSAT thermal infrared spectra. Atmospheric Measurement Techniques, 11,
3815–3828.
Lee, J. D., Mobbs, S. D., Wellpott, A., Allen, G., Bauguitte, S. J. B., Burton, R. R., et al. (2018). Flow rate and source reservoir identification
from airborne chemical sampling of the uncontrolled Elgin platform gas release. Atmospheric Measurement Techniques, 11(3),
1725–1739. https://doi.org/10.5194/amt‐11‐1725‐2018

NISBET ET AL. 45 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Lee, Y.‐Y., Jung, H., Ryu, H.‐W., Oh, K.‐C., Jeon, J.‐M., & Cho, K.‐S. (2018). Seasonal characteristics of odor and methane mitigation and
the bacterial community dynamics in an on‐site biocover at a sanitary landfill. Waste Management, 71, 277–286. https://doi.org/10.1016/
j.wasman.2017.10.037
Leip, A., Skiba, U., Vermeulen, A., & Thomson, R. L. (2018). A complete rethink is needed on how greenhouse gas emissions are quantified
for national reporting. Atmospheric Environment, 174, 237–240.
Leson, G., & Winer, A. M. (1991). Biofiltration: An innovative air pollution control technology for VOC emissions. Journal of the Air &
Waste Management Association, 41, 1045–1054.
Levy, P. E., Burden, A., Cooper, M. D. A., Dinsmore, K. J., Drewer, J., Evans, C., et al. (2012). Methane emissions from soils: Synthesis and
analysis of a large UK data set. Global Change Biology, 18(5), 1657–1669. https://doi.org/10.1111/j.1365-2486.2011.02616.x
Liebetrau, J., Reinelt, T., Agostini, A., & Linke, B. (2018). Methane emissions from biogas plants Methods for measurement, results and
effect on greenhouse gas balance of electricity produced (Book), International Energy Agency, Bioenergy. https://www.ieabioenergy.
com/publications/methane-emissions-from-biogas-plants-methods-for-measurement-results-and-effect-on-greenhouse-gas-balance-of-
electricity-produced/
Liu, Y., Ni, B. J., Sharma, K. R., & Yuan, Z. (2015). Methane emission from sewers. Science of the Total Environment, 524‐525, 40–51.
https://doi.org/10.1016/j.scitotenv.2015.04.029
Liu, Z., Maghirang, R., Murphy, J. P. & DeRouchey, J. (2015). Mitigation of air emissions from Swine buildings through the Photocatalytic
Technology Using UV/TiO2. NPB #13–088, National Pork Board, USA. https://www.pork.org/wp-content/uploads/2015/01/13-088-
LIU-KSt.pdf
Lobert, J. M., Scharffe, D. H., Hao, W. M., & Crutzen, P. J. (1990). Importance of biomass burning in the atmospheric budgets of nitrogen‐
containing gases. Nature, 346(6284), 552–554. https://doi.org/10.1038/346552a0
Lowry, D., Fisher, R.E., France, J.L., Coleman, M., Lanoisellé, M., Zazzeri, G., et al. (2019). Environmental baseline monitoring for shale
gas development in the UK: Identification and geochemical characterisation of local source emissions of methane to atmosphere.
Science of the Total Environment, p.134600.
Lowry, D., Holmes, C. W., Rata, N. D., O'Brien, P., & Nisbet, E. G. (2001). London methane emissions: Use of diurnal changes in con-
centration and δ13C to identify urban sources and verify inventories. Journal of Geophysical Research, 106(D7), 7427–7448. https://doi.
org/10.1029/2000JD900601
Lowry, D. Nisbet, E., Fisher, R., Roddy, A., & O'Brien, P. (2009). Ten years of high‐precision methane isotope data for Mace Head and
London: The influence of Canadian and European sources. 14th WMO/IAEA Meeting of Experts on Carbon Dioxide, Other Greenhouse
Gases and related Tracers Measurement Techniques, GAW Report 186 (WMO TD1487), 44‐47.
Lu, X., Iverach, C. P., Harris, S. J., Fisher, R. E., Lowry, D., France, J. L., et al. (2019). In plume Miller‐Tans time series analyses for improved
isotopic source signature characterisation. EGU General Assembly, Vienna, Austria,7‐12 April 2019. In Geophysical Research Abstracts.
Vol. 21, EGU2019‐11559‐1, 2019.
Luhur, A., Etheridge, D., Loh, Z., Noonan, J., Spencer, J., Day, S. (2018) Characterisation of regional fluxes of methane in the Surat Basin,
Queensland. Final report on Task 3: Broad scale application of methane detection, and Task 4: Methane emissions enhanced modelling.
Report to the Gas Industry Social and Environmental Research Alliance (GISERA). Report No. EP185211, October 2018. CSIRO
Australia. https://gisera.csiro.au/wp-content/uploads/2018/10/GHG-1-Final-Report.pdf
Luo, J., Xu, H., Liu, Y., Chu, W., Jiang, C., & Zhao, X. (2012). A facile approach for the preparation of biomorphic CuO–ZrO2 catalyst for
catalytic combustion of methane. Applied Catalysis A: General, 423‐424(May 2012), 121–129.
Maasakkers, J. D., Jacob, D. J., Sulprizio, M. P., Turner, A. J., Weitz, M., Wirth, T., et al. (2016). Gridded national inventory of US methane
emissions. Environmental Science & Technology, 50(23), 13,123–13,133. https://doi.org/10.1021/acs.est.6b02878
MacDonald, J. A., Skiba, U., Sheppard, L. J., Hargreaves, K. J., Smith, K. A., & Fowler, D. (1996). Soil environmental variables
affecting the flux of methane from a range of forest, moorland and agricultural soils. Biogeochemistry,
34(3). https://doi.org/10.1007/BF00000898
Matsura, N., Hatamoto, M., Sumino, H., Syutsubo, K., Yamaguchi, T., & Ohashi, A. (2015). Recovery and biological oxidatio of dissolved
methane in effluent from UASB treatment of municipal sewage using a two‐stage closed downflow hanging sponge system. Journal of
Environmental Management, 151, 200–209. https://doi.org/10.1016/j.jenvman.2014.12.026
Maurer, D. L., & Koziel, J. A. (2019). On‐farm pilot‐scale testing of black ultraviolet light and photocatalytic coating for mitigation of odor,
odorous VOCs, and greenhouse gases. Chemosphere, 221, 778–784. https://doi.org/10.1016/j.chemosphere.2019.01.086
Mayfield, E. N., Robinson, A. L., & Cohon, J. L. (2017). System‐wide and superemitter policy options for the abatement of methane
emissions from the US natural gas system. Environmental Science & Technology, 51(9), 4772–4780. https://doi.org/10.1021/acs.
est.6b05052
McDermitt, D., Burba, G., Xu, L., Anderson, T., Komissarov, A., Riensche, B., et al. (2011). A new low‐power, open‐path instrument for
measuring methane flux by eddy covariance. Applied Physics B, 102(2), 391–405. https://doi.org/10.1007/s00340-010-4307-0
McKain, K., Down, A., Raciti, S. M., Budney, J., Hutyra, L. R., Floerchinger, C., et al. (2015). Methane emissions from natural gas infra-
structure and use in the urban region of Boston, Massachusetts. Proceedings of the National Academy of Sciences, 112(7), 1941–1946.
https://doi.org/10.1073/pnas.1416261112
Melse, R., & Hol, J. M. G. (2017). Biofiltration of exhaust air from animal houses: Evaluation of removal efficiencies and practical experi-
ences with biobeds at three field sites. Biosystems Engineering, 159, 59–69.
Melse, R. W., & van der Werf, A. W. (2005). Biofiltration for mitigation of methane emission from animal husbandry. Environmental science
& technology, 39, 5460–5468.
Mielke‐Maday, I., Schwietzke, S., Yacovitch, T., Miller, B., Conley, S., Kofler, J., et al. (2019). Methane source attribution in a US dry gas
basin using spatial patterns of ground and airborne ethane and methane measurements. Elementa: Science of the Anthropocene, 7, 13.
https://doi.org/10.1525/elementa.351
Miller, J. B., & Tans, P. P. (2003). Calculating isotopic fractionation from atmospheric measurements at various scales. Tellus Series B:
Chemical and Physical Meteorology, 55, 207–214.
Miller, S. M., Miller, C. E., Commane, R., Chang, R. Y. W., Dinardo, S. J., Henderson, J. M., et al. (2016). A multiyear estimate of methane
fluxes in Alaska from CARVE atmospheric observations. Global Biogeochemical Cycles, 30, 1441–1453. https://doi.org/10.1002/
2016GB005419
Miller, S. M., Michalak, A. M., Detmers, R. G., Hasekamp, O. P., Bruhwiler, L. M. P., & Schwietzke, S. (2019). China's coal mine methane
regulations have not curbed growing emissions. Nature Communications, 10(1), 303. https://doi.org/10.1038/s41467-018-07891-7

NISBET ET AL. 46 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Mitchell, A. L., Tkacik, D. S., Roscioli, J. R., Herndon, S. C., Yacovitch, T. I., Martinez, D. M., et al. (2015). Measurements of methane
emissions from natural gas gathering facilities and processing plants: Measurement results. Environmental Science & Technology, 49,
3219–3227.
Moore, T. A. (2012). Coalbed methane: A review. International Journal of Coal Geology, 101, 36–81. https://doi.org/10.1016/j.
coal.2012.05.011
Naqvi, S. W. A., Lam, P., Narvenkar, G., Sarkar, A., Naik, H., Pratihary, A., et al. (2018). Methane stimulates massive nitrogen loss from
freshwater reservoirs in India. Nature Communications, 9(1), 1265. https://doi.org/10.1038/s41467-018-03607-z
Narsimhan, K., Iyoki, K., Dinh, K., & Román‐Leshkov, Y. (2016). Catalytic oxidation of methane into methanol over copper‐exchanged
zeolites with oxygen at low temperature. ACS Central Science, 2(6), 424–429. https://doi.org/10.1021/acscentsci.6b00139
Nelson, D. D., McManus, B., Urbanski, S., Herndon, S., & Zahniser, M. S. (2004). High precision measurements of atmospheric nitrous
oxide and methane using thermoelectrically cooled mid‐infrared quantum cascade lasers and detectors. Spectrochimica Acta Part A:
Molecular and Biomolecular Spectroscopy, 60(14), 3325–3335. https://doi.org/10.1016/j.saa.2004.01.033
Ni, X., & Groffman, P. M. (2018). Declines in methane uptake in forest soils. Proceedings of the National Academy of Sciences, 115,
8587–8590.
Nisbet, E. G., Dlugokencky, E. J., & Bousquet, P. (2014). Methane on the rise—Again. Science, 343, 493–495.
Nisbet, E. G., Dlugokencky, E. J., Manning, M. R., Lowry, D., Fisher, R. E., France, J. L., et al. (2016). Rising atmospheric methane: 2007–
2014 growth and isotopic shift. Global Biogeochemical Cycles, 30, 1356–1370. https://doi.org/10.1002/2016GB005406
Nisbet, E. G., Manning, M. R., Dlugokencky, E. J., Fisher, R. E., Lowry, D., Michel, S. E., et al. (2019). Very strong atmospheric methane
growth in the 4 years 2014–2017: Implications for the Paris Agreement. Global Biogeochemical Cycles, 33, 318–342. https://doi.org/
10.1029/2018GB006009
Nisbet, E. G., & Weiss, R. (2010). Top‐down vs. Bottom‐up. Science, 328, 1241–1243.
Niu, M., Kebreab, E., Hristov, A. N., Oh, J., Arndt, C., Bannink, A., et al. (2018). Prediction of enteric methane production, yield, and
intensity in dairy cattle using an intercontinental database. Global Change Biology, 24(8), 3368–3389. https://doi.org/10.1111/gcb.14094
NTSB (2018). Preliminary report pipeline: Over‐pressure of a Columbia gas of Massachusetts Low‐pressure Natural Gas Distribution
System. National Transportation Safety Board PLD18MR003‐preliminary‐report. https://www.ntsb.gov/investigations/
AccidentReports/Pages/PLD18MR003-preliminary-report.aspx
Obidzinski, K., Andriani, R., Komarudin, H., & Andrianto, A. (2012). Environmental and social impacts of oil palm plantations and their
implications for biofuel production in Indonesia. Ecology and Society, 17, 25.
O'Brien, D. M., Polonsky, I. N., Utembe, S. R., & Rayner, P. J. (2016). Potential of a geostationary GeoCARB mission to estimate surface
emissions of CO 2, CH 4 and CO in a polluted urban environment: Case study Shanghai. Atmospheric Measurement Techniques, 9(9),
4633–4654. https://doi.org/10.5194/amt-9-4633-2016
O'Connell, E., Risk, D., Atherton, E., Bourlon, E., Fougère, F., Baillie, J., et al. (2019). Methane emissions from contrasting production
regions within Alberta, Canada: Implications under incoming federal methane regulations. Elementa: Science of the Anthropocene, 7, 3.
https://doi.org/10.1525/elementa341
Omara, M., Zimmerman, N., Sullivan, M. R., Li, X., Ellis, A., Cesa, R., et al. (2018). Methane emissions from natural gas production sites in
the United States: Data synthesis and national estimate. Environmental Science & Technology, 52, 12,915–12,925.
O'Shea, S. J., Allen, G., Fleming, Z. L., Bauguitte, S. J. B., Percival, C. J., Gallagher, M. W., et al. (2014). Area fluxes of carbon dioxide,
methane, and carbon monoxide derived from airborne measurements around Greater London: A case study during summer 2012.
Journal of Geophysical Research: Atmospheres, 119, 4940–4952. https://doi.org/10.1002/2013JD021269
O'Shea, S. J., Allen, G., Gallagher, M. W., Bower, K., Illingworth, S. M., Muller, J. B. A., et al. (2014). Methane and carbon dioxide fluxes and
their regional scalability for the European Arctic wetlands during the MAMM project in summer 2012. Atmospheric Chemistry and
Physics, 14(23), 13,159–13,174. https://doi.org/10.5194/acp-14-13159-2014
O'Shea, S. J., Bauguitte, S. J.‐B., Gallagher, M. W., Lowry, D., & Percival, C. J. (2013). Development of a cavity‐enhanced absorption
spectrometer for airborne measurements of CH4 and CO2. Atmospheric Measurement Techniques, 6, 1095–1109. https://doi.org/10.5194/
amt-6-1095-2013
Palmer, P. I., O'Doherty, S., Allen, G., Bower, K., Bösch, H., Chipperfield, M. P., et al. (2018). A measurement‐based verification framework
for UK greenhouse gas emissions: An overview of the Greenhouse gAs Uk and Global Emissions (GAUGE) project. Atmospheric
Chemistry and Physics Discussions, 18(16), 11,753–11,777. https://doi.org/10.5194/acp-18-11753-2018
Pandey, S., Gautam, R., Houweling, S., van der Gon, H. D., Sadavarte, P., Borsdorff, T., et al. (2019). Satellite observations reveal extreme
methane leakage from a natural gas well blowout. Proceedings of the National Academy of Sciences, 116, 26,376–26,381.
Park, S., Lee, C. H., Ryu, C. R., & Sung, K. (2009). Biofiltration for reducing methane emissions from modern sanitary landfills at the low
methane generation stage. Water, Air, and Soil Pollution, 196(1‐4), 19–27. https://doi.org/10.1007/s11270-008-9754-4
Pasquill, F. (1975). Limitations and prospects in estimation of dispersion of pollution on a regional scale. Advances in Geophysics, 18, 1–13.
Pataki, D. E., Bowling, D. R., & Ehleringer, J. R. (2003). Seasonal cycle of carbon dioxide and its isotopic composition in an urban atmo-
sphere: Anthropogenic and biogenic effects. Journal of Geophysical Research, 108(D23), 4735. https://doi.org/10.1029/2003JD003865
Peischl, J., Eilerman, S. J., Neuman, J. A., Aikin, K. C., de Gouw, J., Gilman, J. B., et al. (2018). Quantifying methane and ethane emissions
to the atmosphere from central and western US oil and natural gas production regions. Journal of Geophysical Research: Atmospheres,
123, 7725–7740. https://doi.org/10.1029/2018JD028622
Peischl, J., Ryerson, T. B., Aikin, K. C., De Gouw, J. A., Gilman, J. B., Holloway, J. S., et al. (2015). Quantifying atmospheric methane
emissions from the Haynesville, Fayetteville, and northeastern Marcellus shale gas production regions. Journal of Geophysical Research:
Atmospheres, 120, 2119–2139. https://doi.org/10.1002/2014JD022697
Pétron, G., Frost, G., Miller, B. R., Hirsch, A. I., Montzka, S. A., Karion, A., et al. (2012). Hydrocarbon emissions characterization in the
Colorado Front Range: A pilot study. Journal of Geophysical Research, 117, D04304. https://doi.org/10.1029/2011JD016360
Petrov, A. W., Ferri, D., Krumeich, F., Nachtegaal, M., Van Bokhoven, J. A., & Kröcher, O. (2018). Stable complete methane oxidation over
palladium based zeolite catalysts. Nature Communications, 9(1), 2545. https://doi.org/10.1038/s41467-018-04748-x
Phillips, N. G., Ackley, R., Crosson, E. R., Down, A., Hutyra, L. R., Brondfield, M., et al. (2013). Mapping urban leaks: Methane leaks across
Boston. Environmental Pollution, 173, 1–4. https://doi.org/10.1016/j.envpol.2012.11.003
Pitt, J., Allen, G., Bauguitte, S. J., Gallagher, M. W., Lee, J. D., Drysdale, W., et al. (2019). Assessing London CO2, CH4 and CO emissions using
aircraft measurements and dispersion modelling. Atmospheric Chemistry and Physics, 19, 1–22. https://doi.org/10.5194/acp-2018-1033
Pitt, J. R., le Breton, M., Allen, G., Percival, C. J., Gallagher, M. W., Bauguitte, S. J. B., et al. (2016). The development and evaluation of
airborne in situ N2O and CH4 sampling using a quantum cascade laser absorption spectrometer (QCLAS). Atmospheric Measurement
Techniques, 9(1), 63–77. https://doi.org/10.5194/amt-9-63-2016

NISBET ET AL. 47 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Plant, G., Kort, E. A., Floerchinger, C., Gvakharia, A., Vimont, I., & Sweeney, C. (2019). Large fugitive methane emissions from urban
centers along the US East Coast. Geophysical Research Letters, 46, 8500–8507. https://doi.org/10.1029/2019GL082635
Polag, D., & Keppler, F. (2019). Global methane emissions from the human body: Past, present and future. Atmospheric Environment, 214,
116823. https://doi.org/10.1016/j.atmosenv.2019.116823
Poore, J., & Nemecek, T. (2018). Reducing food's environmental impacts through producers and consumers. Science, 360, 987–992.
Prather, M. (1996). Time scales in atmospheric chemistry: Theory, GWPs for CH4 and CO, and runaway growth. Geophysical Research
Letters, 23, 2597–2600.
Prather, M., Ehhalt, D., Dentener, F., Derwent, R., & Grubler, A. (2001). Atmospheric chemistry and greenhouse gases. In J. T. Houghton,
et al. (Eds.), Climate Change 2001: The Scientific Basis, Contribution of Working Group 1 to the Third Assessment Report of the
Intergovernmental Panel on Climate Change (pp. 239–287). Cambridge, UK: Cambridge Univ. Press.
Pratt, C., & Tate, K. (2018). Mitigating methane: Emerging technologies to combat climate change's second leading contributor.
Environmental Science & Technology, 52, 6084–6097.
Ramankutty, N., Evan, A. T., Monfreda, C., & Foley, J. A. (2008). Farming the planet: 1. Geographic distribution of globalagricultural lands
in the year 2000. Global Biogeochemical Cycles, 22, GB1003. https://doi.org/10.1029/2007GB002952
Ramirez, A. A., García‐Aguilar, B. P., Jones, J. P., & Heitz, M. (2012). Improvement of methane biofiltration by the addition of non‐ionic
surfactants to biofilters packed with inert materials. Process Biochemistry, 47, 76–82.
Ravikumar, A. P., Sreedhara, S., Wang, J., Englander, J., Roda‐Stuart, D., Bell, C., et al. (2019). Single‐blind inter‐comparison of methane
detection technologies–results from the Stanford/EDF Mobile Monitoring Challenge. Elementa: Science of the Anthropocene, 7, 37.
https://doi.org/10.1525/elementa373
Ravikumar, A. P., Wang, J., & Brandt, A. R. (2016). Are optical gas imaging technologies effective for methane leak detection?
Environmental Science & Technology, 51, 718–724.
Reddy, K. R., Yargicoglu, E. N., Yue, D., & Yaghoubi, P. (2014). Enhanced microbial methane oxidation in landfill cover soil amended with
biochar. Journal of Geotechnical and Geoenvironmental Engineering, 140(9), 04014047.
Rees‐White, T. C., Mønster, J., Beaven, R. P., & Scheutz, C. (2018). Measuring methane emissions from a UK landfill using the tracer
dispersion method and the influence of operational and environmental factors. Waste Management, 87, 870–882.
Reinelt, T., & Liebetrau, J. (2019). Monitoring and mitigation of methane emissions from pressure relief valves of a biogas plant. Chemical
Engineering & Technology, 43(1), 7–18. https://doi.org/10.1002/ceat.201900180
Reisinger, A., & Clark, H. (2018). How much do direct livestock emissions actually contribute to global warming? Global Change Biology,
24(4), 1749–1761.
Rella, C. W., Hoffnagle, J., He, Y., & Tajima, S. (2015). Local‐ and regional‐scale measurements of CH4, δ13CH4, and C2H6 in the Uintah
Basin using a mobile stable isotope analyzer. Atmospheric Measurement Techniques, 8, 4539–4559.
Ren, X., Hall, D. L., Vinciguerra, T., Benish, S. E., Stratton, P. R., Ahn, D., et al. (2019). Methane emissions from the Marcellus Shale in
southwestern Pennsylvania and northern West Virginia based on airborne measurements. Journal of Geophysical Research:
Atmospheres, 124, 1862–1878. https://doi.org/10.1029/2018JD029690
Reuters (2018a). Trump's EPA proposes weaker methane rules for oil and gas wells. https://www.reuters.com/article/us-usa-epa-methane/
trumps-epa-proposes-weaker-methane-rules-for-oil-and-gas-wells-idUSKCN1LR2BK
Reuters (2018b). Exxon Mobil opposes weakening Obama‐era emissions rules: Letter to EPA https://www.reuters.com/article/us-exxon-
mobil-epa-methane/exxon-mobil-opposes-weakening-obama-era-emissions-rules-letter-to-epa-idUSKBN1OH23N
Riddick, S. N., Connors, S., Robinson, A. D., Manning, A. J., Jones, P. S. D., Lowry, D., et al. (2017). Estimating the size of a methane
emission point source at different scales: From local to landscape. Atmospheric Chemistry and Physics, 17(12), 7839–7851. https://doi.
org/10.5194/acp-17-7839-2017
Riddick, S. N., Mauzerall, D. L., Celia, M., Harris, N. R. P., Allen, G., Pitt, J., et al. (2019). Measuring methane emissions from oil and gas
platforms in the North Sea. Atmospheric Chemistry and Physics, 19, 9787–9796.
Rigby, M. Montzka, S., Prinn, R., White, J., Young, D., O'Doherty, S., et al. (2017) Hydroxyl radical variability and its influence on recent
methane growth inferred from methyl chloroform trends. European Geoscience Union EGU2017‐10053, 2017
Rigby, M., Montzka, S. A., Prinn, R. G., White, J. W., Young, D., O'Doherty, S., et al. (2017). Role of atmospheric oxidation in recent
methane growth. Proceedings of the National Academy of Sciences, 114(21), 5373–5377.
Rinne, J., Riutta, T., Pihlatie, M., Aurela, M., Haapanala, S., Tuovinen, J. P., et al. (2007). Annual cycle of methane emission from a boreal
fen measured by the eddy covariance technique. Tellus Series B: Chemical and Physical Meteorology, 59(3), 449–457. https://doi.org/
10.1111/j.1600-0889.2007.00261.x
Roberts, H. (2018). Environmental, health and social impacts of dumping and burning of municipal solid waste in South Africa. Linnaeus
Eco‐Tech, 45–45.
Robertson, A. M., Edie, R., Snare, D., Soltis, J., Field, R. A., Burkhart, M. D., et al. (2017). Variation in methane emission rates from well
pads in four oil and gas basins with contrasting production volumes and compositions. Environmental Science & Technology, 51(15),
8832–8840. https://doi.org/10.1021/acs.est.7b00571
Robotics, I. (2017). Robotic and autonomous systems for resilient infrastructure. UK‐RAS White Papers© UK‐RAS.
ISSN 2398‐4422
Rogelj, J., Meinshausen, M., & Knutti, R. (2012). Global warming under old and new scenarios using IPCC climate sensitivity range esti-
mates. Nature Climate Change, 2(4), 248–253. https://doi.org/10.1038/nclimate1385
Rogelj, J., Meinshausen, M., Schaeffer, M., Knutti, R., & Riahi, K. (2015). Impact of short‐lived non‐CO2 mitigation on carbon budgets for
stabilizing global warming. Environmental Research Letters, 10, 075001.
Roscioli, J. R., Yacovitch, T. I., Floerchinger, C., Mitchell, A. L., Tkacik, D. S., Subramanian, R., et al. (2015). Measurements of methane
emissions from natural gas gathering facilities and processing plants: Measurement methods. Atmospheric Measurement Techniques,
8(5), 2017–2035. https://doi.org/10.5194/amt-8-2017-2015
Sadasivam, B. Y., & Reddy, K. (2014). Landfill methane oxidation in soil and bio‐based cover systems: A review. Reviews in Environmental
Science and Biotechnology, 13, 79–107.
Samuelsson, J., Delre, A., Tumlin, S., Hadi, S., Offerle, B., & Scheutz, C. (2018). Optical technologies applied alongside on‐site and remote
approaches for climate gas emission quantification at a wastewater treatment plant. Water Research, 131, 299–309. https://doi.org/
10.1016/j.watres.2017.12.018
Saunois, M., Bousquet, P., Poulter, B., Peregon, A., Ciais, P., Canadell, J. G., et al. (2016). The global methane budget 2000–2012. Earth
System Science Data (Online), 8, 2.

NISBET ET AL. 48 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Saunois, M., Stavert, A. R., Poulter, B., Bousquet, P., Canadell, J. G., Jackson, R. B., et al. (2019). The global methane budget 2000–2017.
Earth System Science Data Discussions (online). https://doi.org/10.5194/essd-2019-128
Schaefer, H., Fletcher, S. E. M., Veidt, C., Lassey, K. R., Brailsford, G. W., Bromley, T. M., et al. (2016). A 21st‐century shift from fossil‐fuel to
biogenic methane emissions indicated by 13CH4. Science, 352(6281), 80–84.
Scheutz, C., & Fredenslund, A. M. (2019). Total methane emission rates and losses from 23 biogas plants. Waste Management, 97, 38–46.
https://doi.org/10.1016/j.wasman.2019.07.029
Scheutz, C., & Kjeldsen, P. (2019). Guidelines for landfill gas emission monitoring using the tracer gas dispersion method. Waste
Management, 85, 351–360.
Scheutz, C., Kjeldsen, P., Bogner, J. E., de Visscher, A., Gebert, J., Hilger, H. A., et al. (2009). Microbial methane oxidation processes and
technologies for mitigation of landfill gas emissions. Waste Management & Research, 27(5), 409–455. https://doi.org/10.1177/
0734242X09339325
Scheutz, C., Pedersen, R. B., Petersen, P. H., Jørgensen, J. H. B., Ucendo, I. M. B., Mønster, J. G., et al. (2014). Mitigation of methane
emission from an old unlined landfill in Klintholm, Denmark using a passive biocover system. Waste Management, 34(7), 1179–1190.
https://doi.org/10.1016/j.wasman.2014.03.015
Scheutz, C., Samuelsson, J., Fredenslund, A. M., & Kjeldsen, P. (2011). Quantification of multiple methane emission sources at landfills
using a double tracer technique. Waste Management, 31(5), 1009–1017. https://doi.org/10.1016/j.wasman.2011.01.015
Schwietzke, S., Harrison, M., Lauderdale, T., Branson, K., Conley, S., George, F. C., et al. (2018). Aerially guided leak detection and repair:
A pilot field study for evaluating the potential of methane emission detection and cost‐effectiveness. Journal of the Air & Waste
Management Association, 69(1), 71–88.
Schwietzke, S., Pétron, G., Conley, S., Pickering, C., Mielke‐Maday, I., Dlugokencky, E. J., et al. (2017). Improved mechanistic under-
standing of natural gas methane emissions from spatially resolved aircraft measurements. Environmental Science & Technology,
2017(51), 7286–7294.
Schwietzke, S., Sherwood, O. A., Bruhwiler, L. M., Miller, J. B., Etiope, G., Dlugokencky, E. J., et al. (2017). Upward revision of global fossil
fuel methane emissions based on isotope database. Nature, 538, 88–91.
Seinfeld, J. H., & Pandis, S. N. (2006). Atmospheric chemistry and physics: From air pollution to climate change. London: John Wiley &
Sons.
Serrano‐Silva, N., Sarria‐Guzmán, Y., Dendooven, L., & Luna‐Guido, M. (2014). Methanogenesis and methanotrophy in soil: A review.
Pedosphere, 24(3), 291–307. https://doi.org/10.1016/S1002-0160(14)60016-3
Shah, A., Allen, G., Pitt, J. R., Ricketts, H., Williams, P. I., Helmore, J., et al. (2019). A near‐field Gaussian plume inversion flux
quantification method, Applied to Unmanned Aerial Vehicle Sampling. Atmosphere, 10(7), 396. https://doi.org/10.3390/
atmos10070396
Sheng, J. X., Jacob, D. J., Maasakkers, J. D., Zhang, Y., & Sulprizio, M. P. (2018). Comparative analysis of low‐Earth orbit (TROPOMI) and
geostationary (GeoCARB, GEO‐CAPE) satellite instruments for constraining methane emissions on fine regional scales: Application to
the Southeast US. Atmospheric Measurement Techniques, 11(12), 6379–6388.
Sheng, J. X., Jacob, D. J., Turner, A. J., Maasakkers, J. D., Benmergui, J., Bloom, A. A., et al. (2018). 2010–2016 methane trends over Canada,
the United States, and Mexico observed by the GOSAT satellite: Contributions from different source sectors. Atmospheric Chemistry and
Physics, 18(16), 12,257–12,267. https://doi.org/10.5194/acp-18-12257-2018
Sherwood, O. A., Schwietzke, S., Arling, V. A., & Etiope, G. (2017). Global inventory of gas geochemistry data from fossil fuel, microbial and
biomass burning sources, version 2017. Earth System Science Data, 9(2). https://doi.org/10.5194/essd-2017-20
Shindell, D. T., Fuglestvedt, J. S., & Collins, W. J. (2017). The social cost of methane: Theory and applications. Faraday Discussions, 200,
429–451. https://doi.org/10.1039/C7FD00009J
Singh, J. S., & Strong, P. J. (2016). Biologically derived fertilizer: A multifaceted bio‐tool in methane mitigation. Ecotoxicology and
Environmental Safety, 124, 267–276.
Smith, K. A., Ball, T., Conen, F., Dobbie, K. E., Massheder, J., & Rey, A. (2018). Landmark papers: No. 7. European Journal of Soil Science,
69, 2–4.
Smith, K. A., et al. (2003). Exchange of greenhouse gases between soil and atmosphere: Interactions of soil physical factors and biological
processes. European J. Soil Science, 54, 779–791.
Smith, M. L., Kort, E. A., Karion, A., Sweeney, C., Herndon, S. C., & Yacovitch, T. I. (2015). Airborne ethane observations in the
Barnett Shale: Quantification of ethane flux and attribution of methane emissions. Environmental Science & Technology, 49,
8158–8166.
Smith, P., Martino, D., Cai, Z., Gwary, D., Janzen, H., Kumar, P., et al. (2008). Greenhouse gas mitigation in agriculture. Philosophical
Transactions of the Royal Society, B: Biological Sciences, 363, 789–813.
Solomon, S., Pierrehumbert, R., Matthews, D., Daniel, J. S., & Friedlingstein, P. (2013). Atmospheric composition, irreversible climate
change, and mitigation policy. In Climate Science for Serving Society: Research, modelling and Prediction priorities (pp. 415–436).
Dordrecht: Springer Science+Business Media. https://doi.org/10.1007/978-94-007-6692-1_15
Sparrow, K. J., Chanton, J. P., Green, R. B., Scheutz, C., Hater, G. R., Wilson, L. C., & Abichou, T. (2019). Stable isotopic determination of
methane oxidation: When smaller scales are better. Waste Management, 97, 82–87.
Stanley, K. M., Grant, A., O'Doherty, S., Young, D., Manning, A. J., Stavert, A. R., et al. (2018). Greenhouse gas measurements from a UK
network of tall towers: Technical description and first results. Atmospheric Measurement Techniques, 11(3), 1437–1458. https://doi.org/
10.5194/amt-11-1437-2018
Stavert, A. R., O'Doherty, S., Stanley, K., Young, D., Manning, A. J., Lunt, M. F., et al. (2018). UK greenhouse gas measurements at two new
tall towers for aiding emissions verification. 12(8), 4495–4518. https://doi.org/10.5194/amt-2018-140
Steiger, J., Bamberger, I., Buchmann, N., & Eugster, W. (2015). Validation of farm‐scale methane emissions using nocturnal boundary layer
budgets. Atmospheric Chemistry and Physics, 15(24), 14,055–14,069. https://doi.org/10.5194/acp-15-14055-2015
Stein, V. B., & Hettiaratchi, P. A. (2001). Methane oxidation in three Alberta oils: Influence of soil parameters and methane flux rates.
Environmetal Technology, 22, 101–122.
Su, S., & Agnew, J. (2006). Catalytic combustion of coal mine ventilation air methane. Fuel, 85(9), 1201–1210.
Subramanian, R., Williams, L. L., Vaughn, T. L., Zimmerle, D., Roscioli, J. R., Herndon, S. C., et al. (2015). Methane emissions from natural
gas compressor stations in the transmission and storage sector: Measurements and comparisons with the EPA greenhouse gas reporting
program protocol. Environmental Science & Technology, 49(5), 3252–3261. https://doi.org/10.1021/es5060258
Susanti, A., & Maryudi, A. (2016). Development narratives, notions of forest crisis, and boom of oil palm plantations in Indonesia. Forest
Policy and Economics, 73, 130–139.

NISBET ET AL. 49 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Syed, R., Saggar, S., Tate, K., Rehm, B. H. A., & Berben, P. (2017). Assessing the performance of floating biofilters for oxidation of methane
from dairy effluent ponds. Journal of Environmental Quality, 46, 272–279.
Thompson, R. L., Stohl, A., Zhou, L. X., Dlugokencky, E., Fukuyama, Y., Tohjima, Y., et al. (2015). Methane emissions in East Asia for 2000‐
2011 estimated using an atmospheric Bayesian inversion. Journal of Geophysical Research: Atmospheres, 120, 4352–4369. https://doi.org/
10.1002/2014JD022394
Townsend‐Small, A., Ferrara, T. W., Lyon, D. R., Fries, A. E., & Lamb, B. K. (2016). Emissions of coalbed and natural gas methane
from abandoned oil and gas wells in the United States. Geophysical Research Letters, 43, 2283–2290. https://doi.org/10.1002/
2015GL067623
Townsend‐Small, A., Marrero, J. E., Lyon, D. R., Simpson, I. J., Meinardi, S., & Blake, D. R. (2015). Integrating source apportionment
tracers into a bottom‐up inventory of methane emissions in the Barnett Shale hydraulic fracturing region. Environmental Science &
Technology, 49(13), 8175–8182. https://doi.org/10.1021/acs.est.5b00057
Townsend‐Small, A., Tyler, S. C., Pataki, D. E., Xu, X., & Christensen, L. E. (2012). Isotopic measurements of atmospheric methane in Los
Angeles, California, USA: Influence of “fugitive” fossil fuel emissions. Journal of Geophysical Research, 117, D07308. https://doi.org/
10.1029/2011JD016826
Turner, A. J., Frankenberg, C., & Kort, E. A. (2019). Interpreting contemporary trends in atmospheric methane. Proceedings of the National
Academy of Sciences, 116, 2805–2813.
Turner, A. J., Frankenberg, C., Wennberg, P. O., & Jacob, D. J. (2017). Ambiguity in the causes for decadal trends in atmospheric methane
and hydroxyl. Proceedings of the National Academy of Sciences of the United States of America, 114(21), 5367–5372. https://doi.org/
10.1073/pnas.1616020114
Turner, A. J., Jacob, D. J., Benmergui, J., Brandman, J., White, L., & Randles, C. A. (2018). Assessing the capability of different satellite
observing configurations to resolve the distribution of methane emissions at kilometer scales. Atmospheric Chemistry and Physics,
18(11), 8265–8278. https://doi.org/10.5194/acp-18-8265-2018
Tyner, D. R., & Johnson, M. R. (2018). A techno‐economic analysis of methane mitigation potential from reported venting at oil production
sites in Alberta. Environmental Science & Technology, 52(21), 12,877–12,885. https://doi.org/10.1021/acs.est.8b01345
UK CCC (2019). Net Zero: The UK's contribution to stopping global warming. Report of the UK Committee on Climate Change, May 2019.
https://www.theccc.org.uk/publications/
UK NAEI (2019). UK National atmospheric emissions inventory. http://naei.beis.gov.uk/overview/pollutants?view=summary-data&pol-
lutant_id=3
UK NIR (2019). UK Greenhouse Gas Inventory 1990 to 2017: Annual report for submission under the Framework Convention on Climate
Change. UK National Inventory Report, Department for Business, Energy & Industrial Strategy. https://uk-air.defra.gov.uk/library/
reports?report_id=981
UNEP (2018). The emissions gap report 2018. United Nations Environment Programme, Nairobi, ISBN 978‐92‐807‐3726‐4. http://www.
unenvironment.org/emissionsgap
UNFCCC (2015). United Nations Framework Convention on Climate Change Paris Agreement https://unfccc.int/process-and-meetings/
the-paris-agreement/the-paris-agreement
Vaillant, S. R., & Gastec, A. S. (1999). Catalytic combustion in a domestic natural gas burner. Catalysis Today, 47, 415.
Van der Heyden, C., Demeyer, P., & Volcke, E. I. P. (2015). Mitigating emissions from pig and poultry housing facilities through air
scrubbers and biofilters: State‐of‐the‐art and perspectives. Biosystems Engineering, 134, 74–93.
Van der Werf, G. R., Randerson, J. T., Giglio, L., Van Leeuwen, T. T., Chen, Y., Rogers, B. M., et al. (2017). Global fire emissions estimates
during 1997‐2016. Earth System Science Data, 9(2), 697–720. https://doi.org/10.5194/essd-9-697-2017
Van der Zaag, A. C., Baldé, H., Crolla, A., Gordon, R. J., Ngwabie, N. M., Wagner‐Riddle, C., et al. (2018). Potential methane emission
reductions for two manure treatment technologies. Environmental Technology, 39, 851–858.
Varon, D. J., Jacob, D. J., McKeever, J., Jervis, D., Durak, B. O. A., Xia, Y., & Huang, Y. (2018). Quantifying methane point sources from
fine‐scale (GHGSat) satellite observations of atmospheric methane plumes. Atmospheric Measurement Techniques, 11(10), 5673–5686.
https://doi.org/10.5194/amt-11-5673-2018
Varon, D. J., McKeever, D. J., Maasakkers, J. D., Pandey, S., Houweling, S., Aben, I., et al. (2019). Satellite discovery of anomalously large
methane point sources from oil/gas production. Geophysical Research Letters, 46, 13,507–13,516. https://doi.org/10.1029/2019GL083798
Vaughn, T. L., Bell, C. S., Pickering, C. K., Schwietzke, S., Heath, G. A., Pétron, G., et al. (2018). Temporal variability largely explains the
top‐down/bottom‐up difference in methane emission estimates from a natural gas production region. Proceedings of the National
Academy of Sciences of the United States of America, 115(46), 11,712–11,717. https://doi.org/10.1073/pnas.1805687115
Vaughn, T. L., Bell, C.S., Yacovitch, T.I., Roscioli, J.R., Herndon, S.C., Conley, S., et al. (2017) Comparing facility‐level methane emission
rate estimates at natural gas gathering and boosting stations. Elementa Science of the Anthropocene 5.NREL/JA‐6A20‐70688.
Veefkind, J. P., Aben, I., McMullan, K., Förster, H., de Vries, J., Otter, G., et al. (2012). TROPOMI on the ESA Sentinel‐5 Precursor: A GMES
mission for global observations of the atmospheric composition for climate, air quality and ozone layer applications. Remote Sensing of
Environment, 120, 70–83. https://doi.org/10.1016/j.rse.2011.09.027
Veltman, K., Rotz, C. A., Chase, L., Cooper, J., Ingraham, P., Izaurralde, R. C., et al. (2018). A quantitative assessment of Beneficial
Management Practices to reduce carbon and reactive nitrogen footprints and phosphorus losses on dairy farms in the US Great Lakes
region. Agricultural Systems, 166, 10–25. https://doi.org/10.1016/j.agsy.2018.07.005
von Fischer, J. C., Cooley, D., Chamberlain, S., Gaylord, A., Griebenow, C. J., Hamburg, S. P., et al. (2017). Rapid, vehicle‐based identifi-
cation of location and magnitude of urban natural gas pipeline leaks. Environmental Science & Technology, 51(7), 4091–4099. https://doi.
org/10.1021/acs.est.6b06095
Wang, J., Xia, F. F., Bai, Y., Fang, C. R., Shen, D. S., & He, R. (2011). Methane oxidation in landfill waste biocover soil: Kinetics and sen-
sitivity to ambient conditions. Waste Management, 31(5), 864–870. https://doi.org/10.1016/j.wasman.2011.01.026
Warner, D., Podesta, S. C., Hatew, B., Klop, G., Van Laar, H., Bannink, A., & Dijkstra, J. (2015). Effect of nitrogen fertilization rate and
regrowth interval of grass herbage on methane emission of zero‐grazing lactating dairy cows. Journal of Dairy Science, 98, 3383–3393.
Wecht, K. J., Jacob, D. J., Sulprizio, M. P., Santoni, G. W., Wofsy, S. C., Parker, R., et al. (2014). Spatially resolving methane emissions in
California: Constraints from the CalNex aircraft campaign and from present (GOSAT, TES) and future (TROPOMI, geostationary)
satellite observations. Atmospheric Chemistry and Physics, 14(15), 8173–8184. https://doi.org/10.5194/acp-14-8173-2014
Weller, Z. D., Roscioli, J. R., Daube, W. C., Lamb, B. K., Ferrara, T. W., Brewer, P. E., & von Fischer, J. C. (2018). Vehicle‐based methane
surveys for finding natural gas leaks and estimating their size: Validation and uncertainty. Environmental Science & Technology, 52,
11,922–11,930.

NISBET ET AL. 50 of 51
19449208, 2020, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019RG000675 by Egyptian National Sti. Network (Enstinet), Wiley Online Library on [11/10/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews of Geophysics 10.1029/2019RG000675

Wennberg, P. O., Mui, W., Wunch, D., Kort, E. A., Blake, D. R., Atlas, E. L., et al. (2012). On the sources of methane to the Los Angeles
atmosphere. Environmental Science & Technology, 46, 9282–9289.
West, O. (1971). Fire, man and wildlife as interacting factors limiting the development of climax vegetation in Rhodesia. Proceedings of the
Annual Tall Timbers Fire Ecology Conference, 11, 121–145. http://talltimbers.org/wp-content/uploads/2014/03/West1971_op.pdf
Whalen, S. C., Reeburgh, W. S., & Sandbeck, K. A. (1990). Rapid methane oxidation in landfill cover soil. Applied and Environmental
Microbiology, 56(11), 3405–3411. https://doi.org/10.1128/AEM.56.11.3405-3411.1990
White, W. H., Anderson, J. A., Blumenthal, D. L., Husar, R. B., Gillani, N. V., Husar, J. D., & Wilson, W. E. (1976). Formation and transport
of secondary air pollutants: Ozone and aerosols in the St. Louis urban plume. Science, 194, 187–189.
Wilkinson, D., Nisbet, E. G., & Ruxton, G. D. (2012). Could methane produced by sauropod dinosaurs have helped drive Mesozoic climate
warmth? Current Biology, 22, R292–R293.
Wofsy, S.C., & Hamburg, S. (2019). MethaneSAT—A New Observing Platform For High Resolution Measurements Of Methane and
Carbon Dioxide. In AGU Fall Meeting 2019. AGU.
Worden, J. R., Bloom, A. A., Pandey, S., Jiang, Z., Worden, H. M., Walker, T. W., et al. (2017). Reduced biomass burning emissions reconcile
conflicting estimates of the post‐2006 atmospheric methane budget. Nature Communications, 8(1), 2227. https://doi.org/10.1038/s41467-
017-02246-0
Wu, J., Li, Q., Chen, J., Lei, Y., Zhang, Q., Yang, F., et al. (2018). Afforestation enhanced soil CH4 uptake rate in subtropical China:
Evidence from carbon stable isotope experiments. Soil Biology and Biochemistry, 118, 199–206. https://doi.org/10.1016/j.
soilbio.2017.12.017
Wunch, D., Jones, D. B. A., Toon, G. C., Deutscher, N. M., Hase, F., Notholt, J., et al. (2019). Emissions of methane in Europe inferred by
total column measurements. Atmospheric Chemistry and Physics, 19(6), 3963–3980. https://doi.org/10.5194/acp-19-3963-2019
Wunch, D., Toon, G. C., Hedelius, J. K., Vizenor, N., Roehl, C. M., Saad, K. M., et al. (2016). Quantifying the loss of processed natural gas
within California's South Coast Air Basin using long‐term measurements of ethane and methane. Atmospheric Chemistry and Physics,
16, 14,091.
Yacovitch, T. I., Neininger, B., Herndon, S. C., van der Gon, H. D., Jonkers, S., Hulskotte, J., et al. (2018). Methane emissions in the
Netherlands: The Groningen field. Elementa: Science of the Anthropocene, 6(1).
Yadav, V., Duren, R., Mueller, K., Verhulst, K. R., Nehrkorn, T., Kim, J., et al. (2019). Spatio‐temporally resolved methane fluxes from the
Los Angeles Megacity. Journal of Geophysical Research: Atmospheres, 124, 5131–5148. https://doi.org/10.1029/2018JD030062
Yang, S., Talbot, R., Frish, M., Golston, L., Aubut, N., Zondlo, M., et al. (2018). Natural gas fugitive leak detection using an unmanned aerial
vehicle: Measurement system description and mass balance approach. Atmosphere, 9(10), 383. https://doi.org/10.3390/atmos9100383
Yargicoglu, E. N., & Reddy, K. R. (2017). Effects of biochar and wood pellets amendments added to landfill cover soil on microbial methane
oxidation: A laboratory column study. Journal of Environmental Management, 193, 19–31.
Yoon, S., Carey, J. N., & Semrau, J. D. (2009). Feasibility of atmospheric methane removal using methanotrophic biotrickling filters.
Applied Microbiology and Biotechnology, 83, 949–956.
Zavala‐Araiza, D., Alvarez, R. A., Lyon, D. R., Allen, D. T., Marchese, A. J., Zimmerle, D. J., & Hamburg, S. P. (2017). Super‐emitters in
natural gas infrastructure are caused by abnormal process conditions. Nature Communications, 8(1), 14012. https://doi.org/10.1038/
ncomms14012
Zavala‐Araiza, D., Lyon, D., Alvarez, R. A., Palacios, V., Harriss, R., Lan, X., et al. (2015). Toward a functional definition of methane super‐
emitters: Application to natural gas production sites. Environmental Science & Technology, 49(13), 8167–8174. https://doi.org/10.1021/
acs.est.5b00133
Zavala‐Araiza, D., Lyon, D. R., Alvarez, R. A., Davis, K. J., Harriss, R., Herndon, S. C., et al. (2015). Reconciling divergent estimates of oil
and gas methane emissions. Proceedings of the National Academy of Sciences, 112(51), 15,597–15,602. https://doi.org/10.1073/
pnas.1522126112
Zazzeri, G., Lowry, D., Fisher, R. E., France, J. L., Lanoisellé, M., Grimmond, C. S. B., & Nisbet, E. G. (2017). Evaluating methane inven-
tories by isotopic analysis in the London region. Scientific Reports, 7(1), 4854. https://doi.org/10.1038/s41598-017-04802
Zazzeri, G., Lowry, D., Fisher, R. E., France, J. L., Lanoisellé, M., Kelly, B. F. J., et al. (2016). Carbon isotopic signature of coal‐derived
methane emissions to the atmosphere: From coalification to alteration. Atmospheric Chemistry and Physics, 16(21), 13,669–13,680.
https://doi.org/10.5194/acp-16-13669-2016
Zazzeri, G., Lowry, D., Fisher, R. E., France, J. L., Lanoisellé, M., & Nisbet, E. G. (2015). Plume mapping and isotopic characterisation of
anthropogenic methane sources. Atmospheric Environment, 110(2015), 151–162. https://doi.org/10.1016/j.atmosenv.2015.03.029
Zhang, B., & Chen, G. Q. ( 2010). Methane emissions by Chinese economy: Inventory and embodiment analysis. Energy Policy, 38,
4304–4316.
Zhang, B., Tian, H., Ren, W., Tao, B., Lu, C., Yang, J., et al. (2016). Methane emissions from global rice fields: Magnitude, spatiotemporal
patterns, and environmental controls. Global Biogeochemical Cycles, 30, 1246–1263. https://doi.org/10.1002/2016GB005381
Zubkova, M., Boschetti, L., Abatzoglou, J. T., & Giglio, L. (2019). Changes in fire activity in Africa from 2002 to 2016 and their potential
drivers. Geophysical Research Letters, 46, 7643–7653. https://doi.org/10.1029/2019GL083469

NISBET ET AL. 51 of 51

You might also like